AN IMPROVED DC-DC CONVERTER FOR PHOTOVOLTAIC POWER SYSTEM APPLICATIONS by Christopher D. Lute A thesis submitted to the Faculty and the Board of Trustees of the Colorado School of Mines in partial fulfillment of the requirements for the degree of Master of Science (Electrical Engineering). Golden, Colorado Date _______________ Signed: _________________________ Christopher Lute Signed: _________________________ Dr. Marcelo G. Simões Thesis Advisor Golden, Colorado Date _______________ Signed: _________________________ Dr. R. Haupt Professor and Head Department of Electrical Engineering and Computer Science ii ABSTRACT The Floating Interleaved Boost Converter (FIBC) is a novel DC-DC converter topology that can provide high voltage gain with improved efficiency and decreased input ripple when compared to conventional converters. These advantages make the FIBC a promising candidate for solar photovoltaic (PV) power system applications. PV modules produce low voltage Direct Current (DC) electricity; this must first be converted to a high voltage DC before being converted to Alternating Current (AC). The FIBC has been demonstrated as a DC-DC converter solution for fuel cells and electric vehicle applications in laboratory settings. In this thesis, the FIBC will be investigated for PV applications, focusing particularly on challenges and solutions for commercialization. This thesis will present a four phase FIBC for PV power system applications. The circuit topology will be analyzed mathematically, and its characteristic advantages will be highlighted. A dual loop, digital linear feedback controller will be developed that is capable of regulating both current and voltage in the device. Additionally, Maximum Power Point Tracking (MPPT), a requisite for PV power converters, will be included and demonstrated. Simulation results from PSIM will be provided to illustrate key findings and control development. The development of a hardware prototype using an embedded microcontroller will be detailed. Finally, the experimental testing and results will be discussed. iii TABLE OF CONTENTS ABSTRACT .............................................................................................................................. iv LIST OF FIGURES ....................................................................................................................... vii LIST OF TABLES .......................................................................................................................... ix LIST OF ABBREVIATIONS .......................................................................................................... x ACKNOWLEDGEMENTS .......................................................................................................... xiii CHAPTER 1 INTRODUCTION ................................................................................................. 1 1.1 Objective ................................................................................................................ 1 1.2 Motivation.............................................................................................................. 2 1.3 Background ............................................................................................................ 2 1.3.1 DC-DC Converter Design Criteria and Characteristics ......................................... 2 1.3.2 DC-DC Boost Converter Topologies ..................................................................... 3 1.3.3 PV Characteristics.................................................................................................. 5 1.3.4 Maximum Power Point Tracking ........................................................................... 7 1.4 PV Power System .................................................................................................. 8 1.5 Scope and Contributions ...................................................................................... 10 1.6 Organization......................................................................................................... 11 CHAPTER 2 CONVERTER ANALYSIS ................................................................................. 13 2.1 Converter Simulation Study................................................................................. 13 2.2 Mathematical Analysis ........................................................................................ 17 2.3 Bidirectional Converter........................................................................................ 28 iv 2.4 Conclusion ........................................................................................................... 29 CHAPTER 3 MAXIMUM POWER POINT TRACKING ....................................................... 31 3.1 MPPT Algorithms ................................................................................................ 31 3.2 Conclusion ........................................................................................................... 35 CHAPTER 4 CONTROL DESIGN ........................................................................................... 36 4.1 Analog Control Design ........................................................................................ 36 4.2 Current Control Loop........................................................................................... 39 4.3 Voltage Control Loop .......................................................................................... 40 4.4 Discrete Controller ............................................................................................... 41 4.5 Conclusion ........................................................................................................... 42 CHAPTER 5 MODELING AND SIMULATION ..................................................................... 44 5.1 Simulation Description ........................................................................................ 44 5.2 Simulation Results ............................................................................................... 45 5.3 Conclusion ........................................................................................................... 48 CHAPTER 6 EXPERIMENTAL PROTOTYPE DEVELOPMENT ........................................ 50 6.1 Inductor ................................................................................................................ 51 6.2 Thermal Management .......................................................................................... 53 6.3 Proof-of-Concept Development ........................................................................... 56 6.4 Snubber Development.......................................................................................... 59 6.5 Gate Driver .......................................................................................................... 62 6.6 Analog Signal Conditioning ................................................................................ 63 6.7 Overcurrent Protection ......................................................................................... 64 v 6.8 Overvoltage Protection ........................................................................................ 66 6.9 Complete Circuit .................................................................................................. 67 6.10 Hardware Design ................................................................................................. 70 6.11 Isolation and Electromagnetic Interference ......................................................... 72 6.12 Microcontroller Code ........................................................................................... 75 6.13 Cost ...................................................................................................................... 77 6.14 Conclusion ........................................................................................................... 77 CHAPTER 7 EXPERIMENTAL TESTING AND RESULTS ................................................. 79 7.1 Current Controller ................................................................................................ 79 7.2 Voltage Controller ............................................................................................... 85 7.3 Maximum Power Point Tracking ......................................................................... 87 7.4 Conclusion ........................................................................................................... 93 CHAPTER 8 CONCLUSION AND FUTURE WORK ............................................................ 94 REFERENCES ............................................................................................................................. 98 APPENDIX A INDUCTOR TEST REPORT ............................................................................ 100 APPENDIX B COMPONENT LIST ......................................................................................... 102 APPENDIX C BILL OF MATERIALS LIST ........................................................................... 113 vi LIST OF FIGURES Figure 1.1 Example of PV pn junction. ........................................................................................................ 5 Figure 1.2 Typical I/V and P characteristics for a PV module under various irradiance values. ................. 7 Figure 1.3 Typical PV power system block diagram. ................................................................................... 9 Figure 1.4 Power outputs for full system simulation. ................................................................................... 9 Figure 1.5 Battery SOC for full system simulation. ................................................................................... 10 Figure 1.6 DC link voltage, inverter voltage and current. .......................................................................... 10 Figure 2.1 Circuit for 4P FIBC. .................................................................................................................. 17 Figure 2.2 4P FIBC with ideal switch representation. ................................................................................ 18 Figure 2.3 4P FIBC, switching state one. ................................................................................................... 18 Figure 2.4 4P FIBC, switching state two. ................................................................................................... 20 Figure 2.5 4P FIBC, switching state three. ................................................................................................. 21 Figure 2.6 4P FIBC, switching state four.................................................................................................... 22 Figure 2.7 4P FIBC, switching state five. ................................................................................................... 23 Figure 2.8 Bidirectional 4P FIBC. .............................................................................................................. 29 Figure 2.9 Power results for bidirectional 4P FIBC. ................................................................................... 29 Figure 3.1 P&O MPPT algorithm. .............................................................................................................. 32 Figure 3.2 IC algorithm flow diagram. ....................................................................................................... 34 Figure 4.1 Open loop input and output voltage response............................................................................ 37 Figure 4.2 Power response for VC and VDC voltage control strategies for a step change in irradiance. .. 37 Figure 4.3 Control block diagram for 4P FIBC. ......................................................................................... 38 Figure 4.4 4P FIBC control loops. .............................................................................................................. 39 Figure 4.5 Bode diagram of the open loop uncompensated and compensated current control loop........... 40 Figure 4.6 Bode diagram of the open loop uncompensated and compensated voltage control loop .......... 41 Figure 5.1 Step response for analog P&O................................................................................................... 45 Figure 5.2 Fixed versus variable delta for IC MPPT, with step change in irradiance ................................ 46 Figure 5.3 Step response for discrete P&O. ................................................................................................ 47 Figure 5.4 Step response for discrete IC. .................................................................................................... 47 Figure 6.1 Completed inductors attached to chassis ................................................................................... 53 Figure 6.2 POC 1. ....................................................................................................................................... 57 Figure 6.3 POC 1 annotated breadboard. .................................................................................................... 57 Figure 6.4 POC 2. ....................................................................................................................................... 58 Figure 6.5 POC 2 annotated breadboard. .................................................................................................... 58 Figure 6.6 VCE (yellow) and IL (blue) waveforms for POC 1, duty cycle = 0.75. .................................... 59 vii Figure 6.7 VDS (yellow) and IL (blue) waveforms for POC 2, duty cycle = 0.75. .................................... 60 Figure 6.8 VDS (yellow) and IL (blue) waveforms for POC 2, duty cycle = 0.75. .................................... 62 Figure 6.9 Gate driver circuit schematic. .................................................................................................... 63 Figure 6.10 CT signal conditioning circuit schematic. ............................................................................... 63 Figure 6.11 Voltage sensor signal conditioning circuit schematic. ............................................................ 64 Figure 6.12 Overcurrent protection circuit schematic................................................................................. 65 Figure 6.13 Overvoltage protection circuit schematic. ............................................................................... 66 Figure 6.14 Complete circuit schematic diagram. ...................................................................................... 68 Figure 6.15 PCB for 4P FIBC. .................................................................................................................... 71 Figure 6.16. Hardware prototype. ............................................................................................................... 72 Figure 6.17 PCB with switching DC-DC converters. ................................................................................. 73 Figure 6.18 Hardware prototype with external linear power supplies. ....................................................... 74 Figure 6.19 Fixed-Point Integer Control Block Diagram ........................................................................... 77 Figure 7.1 Staggered PWM pulses.............................................................................................................. 79 Figure 7.2 Experimental test setup for current control testing. ................................................................... 80 Figure 7.3 Four inductor currents, steady state, 3 A reference. .................................................................. 80 Figure 7.4 Inductor currents IL1 and IL3 and input current IIN, steady state, 3 A reference. ........................ 81 Figure 7.5 Step response for inductor current IL2 for a reference from 4 to 8 A at time t = 2 ms. .............. 82 Figure 7.6 Step response for inductor current IL2 for a reference from 8 to 4 A at time t = 2 ms ............... 82 Figure 7.7 Step response for inductor current IL2 for a reference from 4 to 8 A, multiple cycles. ............. 83 Figure 7.8 Step response for four inductor currents for a reference from 2.5 to 4 A at time t = 1 ms ........ 83 Figure 7.9 Step response for four inductor currents (IL3 not shown) for a reference from 2.5 to 4 A ........ 84 Figure 7.10 Step response for four inductor currents for a reference from 2.5 to 4 A, multiple cycles ..... 84 Figure 7.11 Experimental test setup for voltage control testing. ................................................................ 85 Figure 7.12 Output voltage response to step change in load from 157 Ω to 103 Ω and back .................... 86 Figure 7.13 Input and inductor current response for a step change in load from 157 Ω to 103 Ω ............. 86 Figure 7.14 Four inductor current response for a step change in load from 157 Ω to 103 Ω and back ...... 87 Figure 7.15. MPPT testing setup................................................................................................................. 88 Figure 7.16 Simulation result for IC MPPT method with duty cycle sweep using PV model input. ......... 89 Figure 7.17 Simulation result for IC MPPT method with duty cycle sweep using DC input source ......... 90 Figure 7.18 Input voltage and current response for CV and IC MPPT algorithms in response to ............. 91 Figure 7.19 Power response for CV and IC MPPT algorithms in response to a change ............................ 91 Figure 7.20 Input voltage and current response for CV and IC MPPT algorithms in response to ............. 92 Figure 7.21 Power response for CV and IC MPPT algorithms in response to a change ............................ 92 viii LIST OF TABLES Table 2.1 Device parameters for simulation study...................................................................................... 15 Table 2.2 Device parameters for simulation study...................................................................................... 16 Table 5.1 MPPT Simulation Energy Results Comparison .......................................................................... 48 Table 6.1 Selected Component Ratings and Part Numbers. ....................................................................... 50 Table 7.1 MPPT power comparison. .......................................................................................................... 89 Table B-1 Component List........................................................................................................................ 103 Table C-1 Bill of Materials. ...................................................................................................................... 114 ix LIST OF ABBREVIATIONS 4P A AC ADC AL AT/cm CCM CdTe CIGS Cin cm COSS Cout CS CT CV DC DCR EMI EOFF EON ESR fc FIBC fsw GW hr I IC ID IGBT IIN ISC ISR It,rms J kHz Four phase Amperes Alternating Current Analog to Digital Converter Nominal inductance factor Ampere-turns per centimeter Continuous Conduction Mode Cadmium Telluride Copper Indium Gallium Selenide Input Capacitor centimeter Output capacitance of the MOSFET Output capacitor snubber capacitor Current Transducer Constant Voltage Direct Current DC resistance Electromagnetic Interference turn-off energy turn-on energy Equivalent Series Resistance crossover frequency Floating, Interleaved Boost Converter Switching frequency Gigawatts hour Current Incremental Conductance drain current Integrated Gate Bipolar Transistor Input current Short circuit current Interrupt Service Routine RMS Transistor current Joules kilohertz x Ki KI KP Kv kW L mA MC MHz MIPS MOSFET MPP MPPT ms mV N nH nH/T2 o C/W OCV oz/ft2 P P&O PCB PI PLL PM POC PPE PV PWM rCin rCout rD rDT rL RL RMS RS rT RθCS current sensor gain Integral gain Proportional gain voltage sensor gain Kilowatts Inductor milli-Amperes Motor Control Megahertz Mega-Instructions Per Second Metal Oxide Semiconductor Field Effect Transistor Maximum Power Point Maximum Power Point Tracking millisecond millivolts Initial number of turns nano-Henrys nano-Henrys per Tesla squared degrees Celsius per Watt Open Circuit Voltage ounce per square foot Power Perturb and Observe Printed Circuit Board Proportional-integral Phase Locked-Loop Phase Margin Proof Of Concept Polypropylene Photovoltaic Pulse Width Modulation Input capacitor ESR Output capacitor ESR Diode resistance Switch reverse recovery diode resistance Inductor DC resistance Load resistance Root Mean Square snubber resistor Switch on resistance thermal resistance from case to mounting surface xi RθJC RθSA S SCP SEPIC SOC STC T TA tf THD TJ tr tS tsw TVS ux,avg V VC VCE(on) VDC VDC VDS VDT Vf VGS VOC VOUT Vref Vt,peak W W/m2 Δ η λ µF µH µm µs Ω thermal resistance from junction to case thermal resistance of the heat sink Switching stress Short Current Pulse Single-Ended Primary Inductor Converter State Of Charge Standard Test Conditions Temperature ambient air temperature fall time Total Harmonic Distortion junction temperature rise time Simulation time switching period Transient Voltage Suppression Switch input average value Voltage Capacitor voltage Collector-emitter on voltage Volts Direct Current DC link voltage Drain to source voltage Switch reverse recovery diode threshold voltage Diode forward voltage drop Gate to source voltage Open circuit voltage Output voltage Reference voltage Peak transistor voltage Watts Watts per meter squared MPPT increment/decrement interval efficiency Irradiance micro-Farads micro-Henreys micrometers microseconds Ohms xii ACKNOWLEDGEMENTS I would like to take this time to express my most sincere gratitude for several people whose support, guidance, and insight were indispensable in completing this thesis. First, I must express my appreciation for my advisor, Dr. Marcelo Simões, who offered me the chance to work on this exciting and challenging research project, and who has been helpful and supportive through every step of the process. Working on this has provided a wealth of knowledge that I never would have received in a classroom, and for that I am extremely grateful. I would also like to extend my thanks to the committee, Dr. Ravel Ammerman and Dr. Salman Mohagheghi, for their interest in and support of this research. I gratefully acknowledge The National Science Foundation and The Petroleum Institute, Abu Dhabi, UAE, for their financial support during my study and research. I would also like to thank Robert V. White of Embedded Power Labs for providing insightful design review and suggestions that proved essential to the development of a working prototype. Additionally, Danilo Brandão provided helpful review and guidance on several aspects, and gave me a trove of good Brazilian music to listen to. Magnetics, Inc. of Pittsburgh, PA, generously donated the cores and bobbins for the custom inductors, which is greatly appreciated. Also appreciated is the fine work done by Badger Magnetics of Colorado Springs, CO, who assembled the final inductors. Advanced Circuits in Aurora, CO, did excellent and prompt work on the printed circuit board at a generous discount. Finally, I must thank my wife, Claire, and two children, Matthew and Josephine. They have been an invaluable source of strength and support. xiii CHAPTER 1 INTRODUCTION Renewable energy conversion technologies based on solar and wind sources are very interesting for electricity production because they are non-polluting and do not deplete finite fossil fuel resources. Solar photovoltaic (PV) power systems have found numerous applications, ranging from small, standalone systems to utility-scale, grid-connected power plants [1]. At the end of 2012, the cumulative installed, grid-connected PV capacity in the United States reached approximately 7.4 Gigawatts (GW). Of the nearly 316,000 PV installations which were connected to the grid in 2012, approximately 283,000 (nearly 90%) were residential systems [2]. Solar PV installations tend to be expensive when compared with conventional fossil fuel power generation, but benefit from having no associated fuel cost [3]. However, in order to maximize the return on investment, it is important to maximize the amount of electricity generated by the PV system. This thesis focuses on the development of a novel, effective power converter which will extract the maximum power from a PV array using a Maximum Power Point Tracking (MPPT) control algorithm with a very efficient circuit topology. 1.1 Objective The goal of this thesis is to design, build, demonstrate, and evaluate a novel, direct-current (DC) converter for use in for PV power system applications, with several advantages when compared to the current state-of-art. The system designed in this thesis will serve as part of a residential-scale, or small commercial, PV power system. The interface for the PV modules and inverter will be addressed as part of the design process. A provision for including optional energy storage will be considered as well, although the actual incorporation of energy storage is outside of the scope of this thesis. MPPT capability will be implemented as part of the converter design, in order to maximize energy capture. The design process will include performing mathematical modeling and simulation, with analysis made with simulation software such as PSIM and Matlab-Simulink. After completing the modeling and simulation study, the converter will be implemented in hardware and evaluated. The Floating, Interleaved Boost Converter (FIBC) is a novel topology which has been initially proposed for fuel cell and hybrid electric vehicle applications, where high voltage gain and high efficiency are of great concern [4]. The FIBC has been demonstrated to offer improved voltage boost ratio and efficiency when compared to conventional boost converters [5], [6]. One contribution of this thesis is to adapt this converter for PV applications, and make it suitable for further energy storage integration. An improved, discrete dual-loop linear feedback controller will be developed and implemented. Different 1 MPPT techniques will be evaluated and demonstrated. A hardware prototype will be developed and experimental results presented. This prototype will exhibit a significantly higher fidelity than previous laboratory prototypes of this converter topology by more closely representing commercially produced hardware. It will use an embedded microcontroller to execute all control functions using fixed-point arithmetic. The prototype will also incorporate several ancillary circuit features and functions, including: analog signal measurement and conditioning, gate drivers, and overcurrent and overvoltage protection. The final prototype will be much closer to a commercial product, or an industrial retrofit, than the prototypes already developed by Kabalo, et al in [5] or Choi, et al in [7]. Specific challenges and unique characteristics will be identified as they relate to commercialization. A bidirectional version of the converter for battery charging applications will be presented and demonstrated in simulation. However, the hardware implemented in this thesis is unidirectional; a bidirectional prototype for battery storage is suggested as a future effort. 1.2 Motivation Recent decreases in the cost of solar technologies have spurred wider deployment of PV electricity generation systems. Two studies produced by the US Department of Energy, the SunShot Vision [8], and the Renewable Energy Futures study [9], both anticipate continued, significant growth in PV between now and the coming decades. These two studies estimated growth in solar capacity (including both PV and concentrated solar power), and the proportion of energy generated by solar technologies up to the year 2050. There was a wide range in results, depending primarily on cost trajectories, but also influenced by grid flexibility and transmission constraints. The study results estimate the installed PV capacity in 2050 to range anywhere from 100 GW to > 600 GW, with decreasing installed cost being the primary driver [10]. Meeting these goals of increased installation at reduced cost will require advanced, high performance power converters capable of maximizing energy conversion. 1.3 Background This section provides background information and literature review on: DC-DC converter design criteria and characteristics, DC-DC boost converter topologies, PV characteristics, and MPPT. 1.3.1 DC-DC Converter Design Criteria and Characteristics An important design criterion for DC-DC converters for PV applications is the ability to produce a high voltage output from a low voltage input. The operating voltage for commercially available PV modules is low voltage dc, typically around 20 volts direct current (VDC) [11]. However, a high input voltage is necessary for efficient conversion to alternating current (AC) when using a DC-AC inverter. 2 One option is to increase this voltage through series combination of multiple PV modules. However, this approach suffers from several disadvantages. First, series combinations reduce reliability; a failure of any one module in a series string will result in the entire string being unavailable. As the current through series-connected components must always be equal, the total string current will be dictated by the lowest performing module in the string. This is especially problematic in partial-shading situations, where shading of a single module will diminish the power output of the entire string. Therefore, it is advantageous to combine modules in parallel and use a DC-DC converter to increase the output voltage to the required input of the DC-AC inverter. The output voltage of this DC-DC converter must be in the range of 400 VDC for a typical DC link input voltage for a three-phase 208/120 V or single-phase 220 V or 240 V inverter. Other important criteria for the design include efficiency and cost. As with any renewable energy application, efficiency is of paramount importance. Cost is another important factor in converter design. Therefore, the converter should employ the least complex topology necessary to achieve the stated goals, and should employ methods to use less expensive components when possible. In addition, the converter may need to address criteria related to ripple and isolation. Ripple refers to the deviation of the input and output voltage and current from a pure DC. Ripple occurs in DC-DC converters due to the switching transients. High input current and voltage ripple can negatively impact the amount of power produced by the PV array, and may interfere with the proper operation of MPPT algorithms. The other factor to consider is isolation; DC-DC converters may be non-isolated or isolated. Galvanic isolation is accomplished through the use of a high frequency transformer in the DC-DC converter, or by a low frequency output transformer. However, these transformers add to the size, weight, and cost of converters, and have associated losses. Therefore, galvanic isolation is typically neglected for PV power converters systems [12]. 1.3.2 DC-DC Boost Converter Topologies Many different DC-DC converters exist that can produce an output voltage that is greater than the input voltage. Conventional boost converter topologies include: boost, buck-boost, Cúk, Single Ended Primary Inductor Converter (SEPIC), and Zeta converters. A DC-DC boost converter produces an output voltage which is always greater in magnitude than the input voltage [13]. The practically realizable voltage gain of this converter in most cases is not greater than 4, particularly when high output voltage and power are required, making this impractical for this application [7]. The boost converter works by closing a switch which isolates the output stage while charging an inductor. When the switch is open, energy from both the input and inductor flow to charge an output capacitor [14]. For buck-boost 3 converters, the magnitude of the output voltage can be greater or less than the input voltage, and has opposite polarity. Notable disadvantages of this circuit include: high input-voltage ripple and high switching stress [13]. Like the buck-boost converter, the Cúk produces an inverter output which can be greater or less than the input voltage in magnitude. This converter has lower input and output current ripple than other converter topologies [14], and has been used for many PV power system applications [13]. The SEPIC can, like the buck-boost and Cúk converters, produce an output voltage which can be either greater or less than the input, but without inverting. Finally, the Zeta converter can also provide a non-inverter output which can be greater or less than the input in magnitude. The voltage gain which can be achieved using any DC-DC converter topology is limited by the parasitics of the components [7]. As power, and current, through a device increases, the conduction (I2R) losses will have greater impacts. Therefore, new high-gain converter topologies are required to provide the necessary voltage gain at high power. The FIBC consists of two conventional boost converters, a floating version on the positive rail and a non-floating version on the negative rail. These are connected in series at the output with the input source. This series connection increases the obtainable voltage gain while reducing the output voltage ripple [7]. The inductor branch can then be split into multiple parallel phases, which reduces the amount of current through each branch, thus reducing the loss. Interleaving converters have reduced current ripple, but the classical interleaved converter does not provide improved voltage gain over conventional boost converters [7]. By combining two boost converters together, and including interleaving, the FIBC can produce high voltage gain with lower current ripple. Although the FIBC requires a greater number of components than conventional converters, it benefits from: Increased efficiency and voltage gain [5] Decreased input ripple [5] Lower voltage and current ratings for semiconductor devices (e.g., switches, diodes) [5] Higher switch utilization factors and lower switching stress [5] Smaller inductor volumes due to lower current flow through each inductor [7] Lower capacitor voltage ratings, as each capacitor carries only about 60% of the output voltage [7] Lower I2R losses and voltage drop through inductive components [7] 4 These advantages have led the FIBC topology to be proposed as a suitable solution for fuel cell and hybrid electric vehicle applications [4]. It has also been proposed as a promising candidate for PV power systems, but has not yet been demonstrated for this particular application [7]. Open loop hardware demonstrations of the FIBC have been conducted in [6] and [7]. A laboratory prototype was demonstrated for fuel cell applications by Kabalo, et al., in [5] and [15]. This prototype made use of a dual-loop, sliding mode controller to regulate inductor current and output voltage [16]. This prototype used a dSPACE DS 1104 real time board for control implementation. The FIBC has also been demonstrated for electric vehicle applications using a linear feedback control algorithm implemented by a digital signal controller [17]. 1.3.3 PV Characteristics PV technology is capable of converting sunlight directly into electricity. The building block of PV technology is the PV cell, which is a specially designed pn junction (or multiple junctions for multilayer PV cells) using a semiconductor material. An example of a pn junction is provided in Figure 1.1. Single crystal (mono-crystalline) silicon tends to be the most robust, efficient semiconductor material for producing PV cells. Unfortunately, producing mono-crystalline silicon tends to be relatively energy intensive, making them more expensive. Multi-crystalline (poly-crystalline) silicon PV cells are more economical, but are typically less efficient. Newer, thin-film semiconductor technologies have been introduced. These materials, which include amorphous silicon, Copper Indium Gallium Selenide (CIGS), and Cadmium Telluride (CdTe) are unique in that only a very thin layer, 1-2 micrometers (µm) is required, which reduces material costs [11]. Figure 1.1 Example of PV pn junction. A typical PV cell will yield < 3 Watts (W) at around 0.5 VDC. In order to produce more useful power and voltage, multiple cells are combined into modules. Each module consists of around 36 cells, 5 combined in series/parallel combinations to obtain the desired voltage and power rating. Module power output ranges from a few watts to > 300 W. Open circuit voltage ratings of around 20-30 VDC are common. Cells are mounted and interconnected with wires or metal foil strips, then encapsulated with a protective glass or composite. Anticipated module lifetime is normally over 20 years [11]. Most applications require more power than is available from a single module. Therefore, multiple modules are combined in series/parallel arrangements into arrays. As has been mentioned previously, series combination will increase the array voltage, but comes with performance and reliability penalties. Parallel combination of PV modules requires each module to be maintained at the same voltage. This will be accomplished by the DC-DC converter, which will determine the operating voltage of the PV array based on the MPPT algorithm. One disadvantage to this method is that the MPPT will be applied for the array as a whole, and cannot take into account variations in individual modules or partial shading conditions. Module-integrated converters have been proposed as a solution for this problem. Every panel would have an integrated DC-DC converter with MPPT to bias each module individually [17]. The current-voltage (I-V) characteristic of PV devices is unique in that it can be approximated as an ideal current source for part of its operational range, and as an ideal voltage source for the other part. In addition, this I-V characteristic is dependent on the operating irradiance (λ) and temperature (T). The short circuit current (ISC) is strongly dependent on λ, but is relatively insensitive to changes in T. Cell open circuit voltage (VOC) decreases with increasing temperature by approximately 0.5%/ oC above rated. This decreased voltage also limits the power produced by a PV cell at higher operating temperature [11]. Typical I-V relationships for a PV module under various irradiance values are given in Figure 1.2. Irradiance is given in watts per meter squared (W/m2). This graph also plots power output (P) for the module. From this, is can be observed that there is a point at which the power produced by the PV module is at its maximum value; this corresponds to the knee point of the I-V curve. An important capability for any effective PV application is the ability to dynamically track this value under varying operating conditions. This task is accomplished through MPPT. 6 10 160 9 140 1,000 W/m2 8 Current [A] 100 6 600 W/m2 5 80 4 60 3 40 200 W/m2 2 1 20 0 0 0 5 Power [W] 120 7 10 15 20 25 Voltage [V] 30 35 Current Power Figure 1.2 Typical I/V and P characteristics for a PV module under various irradiance values. 1.3.4 Maximum Power Point Tracking Extracting the maximum useful power from renewable energy systems, including PV, is essential for maintaining cost effectiveness. In PV applications, this is accomplished through the use of a MPPT control scheme. At least nineteen distinct MPPT methods have been developed in the past [20]. Simple algorithms include Constant Voltage (CV), in which the converter maintains the PV array at a constant reference voltage value (Vref) that has been determined to yield the maximum power under a certain set of conditions. This is therefore an approximate MPPT, and cannot achieve the true maximum power point (MPP). It is very simple and inexpensive to implement, but is not capable of adapting to dynamically changing operating conditions. However, the CV method has been proven to be more effective than other, more complex algorithms in low irradiance conditions; for this reason, CV can be combined with other MPPT techniques [21]. Related to the CV method is the Open Circuit Voltage (OCV) method, which periodically open circuits the PV array. A PV output voltage set point is determined as a certain percentage of VOC, typically between 71-78% [20]. This algorithm is capable of compensating for temperature effects, which affect the output voltage of the PV. A downside is that no power is generated while the PV array is in open circuit. Again, this method can only approximate the MPP. Another simple method for MPPT is the Short Current Pulse (SCP) method. This is akin to the OCV method, but the panel is short-circuited rather than open-circuited. The converter is commanded by a current control loop and the operating current is commanded to be a percentage of the ISC, often around 7 92% [21]. Like the OCV method, no power is produced while the PV array is being short circuited. Because ISC is dependent on irradiance, but is relatively insensitive to temperature, the SCP method is more effective at responding to varying irradiance than to differences in module temperature. Also, the converter will experience large stresses during the switching operations; this is true for both OCV and SCP. Like the CV and OCV, the SCP method only approximates the MPP. More complex methods are available which are capable of determining the true MPP. Extremumseeking control theory can be used to establish a feedback system in order to induce oscillations around the equilibrium point, thereby attempting to determine the MPP [22]. The Perturb-and-Observe (P&O) method periodically adjusts the operating point and measures the instantaneous power output. If the power increases, the converter will continue to adjust the operating point in the same direction, and if it decreases the direction of adjustment will be reversed. This method is effective but has a tendency to oscillate around the MPP and may not be able to adjust to rapidly varying conditions [21]. The Incremental Conductance (IC) method improves on the P&O method be eliminating the oscillations around the output. This method uses the relationship between the instantaneous and incremental conductance (I/V and ) values to determine both the magnitude and direction of adjustment required to achieve the MPP. It requires more complex controls and greater computational resources [21]. Both P&O and IC algorithms may make use of Fuzzy Logic to increase performance and accuracy [20], [22]. In [21], a hybrid combination of IC and CV was found to yield the highest energy capture in a study of 10 different MPPT techniques [21]. 1.4 PV Power System A typical PV power system block diagram is shown in Figure 1.3. Here, a PV array is connected to a DC-DC boost converter to raise it to the required DC link voltage. An energy storage system (e.g., battery) connects through a bidirectional charge controller that steps the battery voltage up to the DC link, and allows power to flow either from the DC link to the battery (charging) or from the battery to the DC link (discharging). An inverter is connected to the DC link in order to produce AC power, with an LCL filter to limit the harmonics [24]. For grid-connected systems, this AC output would be connected to the utility grid, along with any AC loads. The inverter could be unidirectional or bidirectional, depending on whether it is desired to allow grid power to charge the battery. 8 Figure 1.3 Typical PV power system block diagram. A full day simulation of this system was conducted using PSIM, a power electronics simulation software. One simulation second was equal to eight hours of real time. At the start, the PV was producing close to full rated output, supplying the 2 kW resistive load, and charging the battery. At simulation time (tS) = 8 hours (hr), the PV stopped producing power, and the battery supplied the 2 kW load. At tS ≈ 19 hr, the battery state of charge (SOC) dropped to about 10%. At this point, the battery disconnected, and the load was supplied by the grid. The power outputs from the various sources are plotted in Figure 1.4. Figure 1.4 Power outputs for full system simulation. Figure 1.5 shows the battery SOC, and Figure 1.6 plots the DC link voltage, and the inverter voltage and current waveforms. 9 Figure 1.5 Battery SOC for full system simulation. Figure 1.6 DC link voltage, inverter voltage and current. These results illustrate how different power electronics elements interact to supply the load under various operating conditions. Here, it was not desired to use grid power to charge the battery, or to inject surplus power into the grid, thus the grid power was zero until neither the PV nor battery were available, at which point it was used to serve the load. The waveforms illustrate the effectiveness of the LCL filter in providing a sinusoidal voltage and current. Total Harmonic Distortion (THD) was determined to be around 4.9% for the current, and was negligible for the voltage waveform. The DC link voltage was maintained around 400 V with minimal ripple. 1.5 Scope and Contributions The contributions of this these include: 1) performing a quantitative comparison of DC-DC converter performance, 2) developing an improved, discrete, dual-loop, linear feedback controller for the 10 FIBC, 3) implementing an advanced MPPT capability, and 4) development of a high fidelity hardware prototype and presenting experimental results. Many DC-DC converter topologies have been implemented, from very simple boost converter circuits to very complex resonant mode converters. Chapter 2 includes a quantitative simulation analysis of FIBC compared to other, more conventional converter topologies. Each will be evaluated using a simulated PV array, at rated load, with simple open-loop controls dictating the switching duty cycle. In order to have a more realistic hardware implementation of the FIBC, an embedded microcontroller was chosen for implementation, rather than a real time simulation engine. This requires the development of a discrete control algorithm that can be executed using a commercially available microcontroller. This will use a linear feedback control architecture and will control both the inductor currents and output capacitor voltagse. MPPT control will also be included. The performance of different MPPT techniques in both the analog and discrete domains will be evaluated, and one method will be chosen for implementation. The previous hardware prototypes of the FIBC topology as described in [6], [7] were operating in open-loop mode. It was also demonstrated for fuel cell applications using a dSPACE real time simulator [5], [15], [16]. A low fidelity laboratory prototype using an embedded microcontroller was demonstrated for electric vehicle applications [17]. This high fidelity prototype will designed specifically to demonstrate the FIBC for PV system applications. It will incorporate MPPT capability. The prototype will be a closer approximation of commercial hardware. All control functions will be executed by an embedded microcontroller. Additionally, the prototype will include additional ancillary circuits that are necessary for commercial hardware, including: gate drivers, analog signal measurement, and overvoltage and overcurrent protection. This prototype will be demonstrated and experimental results presented. 1.6 Organization Chapter 1 is an introduction, with key motivation and background information, highlighting the research and project goals. Chapter 2 provides additional background on DC-DC converters in general, and the FIBC in particular, with a complete mathematical analysis coverage. Chapter 3 conducts an investigation of MPPT algorithms and techniques and a description of the MPPT algorithm to be implemented in the FIBC. Chapter 4 describes the controller development and implementation. 11 Chapter 5 presents results of simulation studies which will confirm the correct operation and will support the definition of performance parameters. Chapter 6 details the design of the hardware prototype and provides justification for various design decisions. Chapter 7 describes the experimental setup and presents the results of the hardware testing. Chapter 8 contains the final discussion, conclusion, and opportunities for future work. 12 CHAPTER 2 CONVERTER ANALYSIS This chapter presents a quantitative comparison of the FIBC with conventional DC-DC boost converter topologies. Converter performance was evaluated using different criteria, such as: efficiency, voltage/current ripple, switching stress, and switch utilization. A mathematical analysis of the four phase FIBC is presented, including development of the state space equations, static transfer function, and openloop transfer functions for inductor current and capacitor voltage. 2.1 Converter Simulation Study A simulation study of several different DC-DC converter topologies was conducted. Several converter characteristics were calculated and compared to ascertain the most appropriate converter topology for implementation. Efficiency is an important consideration with any device design or implementation. In renewable energy applications, where the cost for energy may be higher than conventionally produced electricity, minimizing loss is of even greater importance. Therefore, the efficiency of each converter under rated conditions (Vout = 400 V, P = 5 kW) was evaluated. Efficiency (denoted by η) was calculated by the ratio of input to output power, as shown in the following equation. (2.1) Ripple refers to the variation between a signal (either voltage or current) and its average value. Excessive ripple may impair system performance by interfering with controls, or may disrupt downstream devices which require a more constant voltage and/or current. Ripple can be minimized by appropriate filtering, using passive elements such as inductors and capacitors. Large passive filter elements, however, will add to the size and cost of the converter, and present a risk of component failure. The energy storage aspect of passive filter elements will also cause the device to respond more slowly to changes in operating conditions and control signals. Ripple is calculated by the following equation: ( ) ( ) (2.2) 13 Although this equation uses voltages, the same equation may be applied to current for determining ripple. Ripple was calculated for input voltage and current, output voltage and current, and inductor current. The maximum, minimum, and average values of each signal were determined while operating under steady state conditions for a 100 millisecond (ms) duration. Inductor current was an important criterion that was evaluated for each converter topology. In this case, the ratio of inductor current to input current was compared. Large inductors with high current carrying capability are expensive and bulky. Therefore, converters with lower average inductor currents are preferable. Switching stress refers to the deleterious effects of high voltages and currents on switching elements, either typically Integrated Gate Bipolar Transistors (IGBTs) or Metal Oxide Semiconductor Field Effect Transistors (MOSFETs). Higher stresses require a more robust device capable of sustaining the higher currents and/or voltages. Devices subjected to increased stress are also at higher risk for failure. Switching stress was calculated as the product of the root mean square (RMS) transistor current (It,rms) and the peak transistor voltage (Vt,peak), and is expressed in W, as shown in the equation below. ( ) ( ) ( ) (2.3) Switch utilization is defined as the ratio of switching stress to output power. Therefore, a converter with higher switch utilization is able to produce a greater power output for a given switching stress. This is illustrated in the following equation. (2.4) A total of eight different DC-DC converter topologies were simulated and analyzed. All topologies were non-isolated converter types. Isolated converters require transformers which add to a converters cost, complexity, size, and weight. Additionally, transformers add to the total converter losses, contributing to lower overall efficiency. For these reasons, non-isolated converters are typically used for PV power system applications [12]. 14 In order to maintain an equal basis for comparison, device ratings and specifications were kept as similar as possible between different converter topologies. Table 2.1 lists these device ratings and parameters used for the converter simulation study. Table 2.1 Device parameters for simulation study. Device Description Rating Cin Input capacitor 300 µF rCin Input capacitor Equivalent Series Resistance (ESR) 80 mΩ L Inductor 30 µH rL Inductor DC resistance 40 mΩ rT Switch on resistance 4.2 mΩ VDT Switch reverse recovery diode threshold voltage 2V rDT Switch reverse recovery diode resistance 11.4 mΩ fsw Switching frequency 19.9 kHz Vf Diode forward voltage drop 2V rD Diode resistance 11.4 mΩ Cout Output capacitor 1,000 µF rCout Output capacitor ESR 120 mΩ RL Load resistance 32 mΩ The results for the six converters are summarized in Table 2.2. The PV was operating under Standard Test Conditions (STC), defined as an irradiance of 1,000 W/m2 and temperature of 25oC. The output voltage (VOUT) was 400 VDC and the output power was 5 kW. 15 Table 2.2 Device parameters for simulation study. Converter η VIN Ripple VOUT Ripple IIN Ripple IOUT Ripple Switching Stress (kW) Switch Utilization 1.0 37.156 0.13 1.121 46.225 0.09 IL,avg /IIN,avg Boost 91.0% 9.5% 4.1% 74.8% 1,105.3% BuckBoost 86.0% 21.7% 4.6% 148.8% 4.6% Cúk 84.8% 10.0% 2.4% 72.8% 2.4% 1.0 49.695 0.09 SEPIC 83.1% 9.7% 5.8% 75.2% 5.8% 1.0 50.863 0.08 Zeta 84.0% 27.7% 3.1% 176.4% 3.1% 1.0 48.157 0.09 2P FIBC 90.1% 6.4% 3.5% 47.3% 3.5% 0.561 12.023 0.39 4P FIBC 94.4% 1.5% 0.5% 11.5% 0.5% 0.282 6.728 0.73 6P FIBC 94.9% 0.6% 0.3% 4% 0.2% 0.187 5.009 0.99 Of the eight converters which were simulated, only four were at least 90% efficient under rated conditions: the boost converter and the three different FIBC topologies. These included the two phase (2P), four phase (4P), and six phase (6P) FIBC. Of these three, the 6P FIBC had the highest efficiency; however, it was only marginally higher than the 4P efficiency. FIBC topologies had the lowest input current ripple among any of the topologies which were investigated. Because the interleaved topology splits the current among multiple inductors, the average current through each inductor was much lower. The boost converter had very high output current ripple, which would require a large low pass filter. The FIBCs had much lower output current ripple, and lower output voltage ripple, than the other converter types. Because of the fact that they use multiple switching devices, the FIBCs had much lower switching stress and higher switch utilization than any other topology. Thus, while they do require more switching devices, FIBCs would be able to use less expensive switches with lower voltage and current specifications, and these devices would be subjected to lower stress. 16 2.2 Mathematical Analysis First, the average state space equations governing the 4P FIBC were developed, using the methodology [18]. Note that all equations developed in this section are relevant only when the converter is operating in continuous conduction mode (CCM). The circuit for the 4P FIBC is shown in Figure 2.1. Figure 2.1 Circuit for 4P FIBC. The equations were derived assuming ideal devices and components. Figure 2.2 shows the ideal 4P FIBC with ideal switching elements. The switching states are defined such that ux = 0 corresponds to the non-conducting mode of switch x (where x = 1, 2, 3, or 4); the corresponding diode dx will be conducting. Therefore, a total of 24 (16) states are possible. 17 Figure 2.2 4P FIBC with ideal switch representation. The first state examined is one in which all four switches are non-conducting (u1 = u2 = u3 = u4 = 0). This is shown in Figure 2.3. Figure 2.3 4P FIBC, switching state one. 18 The six state variables for this state can be expressed as: (2.5) (2.6) (2.7) (2.8) ⁄ (2.9) ⁄ (2.10) (2.11) where the output DC link voltage (vDC) can be expressed in terms of the input voltage and state variables by the equation: (2.12) This expression for the output voltage is valid for all switching states. The next state examined is u1 = u2 = u3 = 0, u4 = 1. This state is illustrated in Figure 2.4. 19 Figure 2.4 4P FIBC, switching state two. The state equations for this state were determined to be: (2.13) (2.14) (2.15) (2.16) ⁄ ⁄ 20 (2.17) (2.18) The third state is defined as u1 = u2 = u4 = 0, u3 = 1. The circuit corresponding to this state is shown in Figure 2.5, followed by the state equations for this state. Figure 2.5 4P FIBC, switching state three. (2.19) (2.20) (2.21) (2.22) ⁄ ⁄ 21 (2.23) (2.24) The fourth state is u1 = u3 = u4 = 0, u2 = 1. This state is illustrated in Figure 2.6. Figure 2.6 4P FIBC, switching state four. The state equations for this state are: (2.25) (2.26) (2.27) (2.28) (2.29) ⁄ ⁄ 22 (2.30) The fifth, and final, state which was examined is u2 = u3 = u4 = 0, u1 = 1, and is illustrated in Figure 2.7, followed by the corresponding state equations. Figure 2.7 4P FIBC, switching state five. (2.31) (2.32) (2.33) (2.34) (2.35) ⁄ ⁄ 23 (2.36) State-space equations can then be developed for each switching state of the converter. Using an average value for each switch (ux,avg), the average state space equations can be written as: ( ) ( ) (2.38) ( ) (2.39) ( ) (2.40) (2.37) ( ) ( ) ⁄ (2.41) ( ) ( ) ⁄ (2.42) Next, it is assumed that all inductors and capacitors have equal values, that is L1 = L2 = L3 = L4 = L, and C1 = C2 = C. The average values of each switch are also assumed to be equal, thus u1,avg = u2,avg = u3,avg = u4,avg = U. The state equations can then be rewritten as normalized average equations by making the following substitutions [18]: 24 √ √ √ √ ( ) ( √ ( (2.43) ) ) The six average, normalized state space equations for the 4P FIBC can be expressed as: ̇ (2.44) ̇ (2.45) ̇ (2.46) ̇ (2.47) (2.48) ̇ (2.49) ̇ where the normalized output voltage is and 25 √ . The control objective is to regulate the output of the converter to an equilibrium value. This is done by applying the appropriate control input u which results in an average switch value U. The 4P FIBC is typically operated with all switches operating with the same duty cycle, each phase shifted 90 o apart [15]. Therefore, for normal operation, the average values of the four switches u1,avg = u2,avg = u3,avg = u4,avg = U. Under normal, steady state operation where the output variable(s) are controlled to their equilibrium values, the time derivatives of the state variables will be zero. Therefore, the static transfer function of the converter can be determined by setting the normalized average state equations to zero, using the average equilibrium values of the state currents and voltages (written as ̅̅̅) [18]. ( ) ( ( ) ̅̅̅ ̅̅̅ ) ̅̅̅ ̅̅̅ ̅̅̅ (̅̅̅) ) ( ( ) ( ) ( ( ) ( ) (2.50) ( ) ) Solving these equations yields: ̅̅̅ ( ̅̅̅ ̅̅̅ ) ( ̅̅̅ ) ̅̅̅ ( ) (2.51) (2.52) ( ) (2.53) ̅̅̅ Assuming that the steady state, equilibrium current is identical through each inductor, the first two equations can be rewritten as: ̅̅̅ ̅̅̅ ̅̅̅ ̅̅̅ ( ) ( ) The steady state, static transfer function for the output voltage is thus determined to be: 26 (2.54) ( ) ̅̅̅ ̅̅̅ (2.55) This reveals one of the implicit advantages of the FIBC as compared to a conventional boost converter, whose static transfer function is [14]: ( ) (2.56) The (1+U) expression in the numerator of the static transfer function for the FIBC allows it to produce a higher voltage gain for the same duty cycle than a conventional boost converter. Next, it is assumed that in each switching period the average voltage across the inductors and the average current through the capacitors are null. In addition, in steady state the average current is assumed to be identical through each inductor, and the average voltage is equal across both capacitors. Then, the output/input voltage ratio (voltage gain) can be found by analyzing one of the passive elements that transfers energy from input to output. Choosing the inductor L1 operating in CCM, the following equation can be defined: ( )[ ( )] (2.57) Based on (54) and (56) the desired transfer functions may be found, such that U’ is the duty cycle complement (1-U) [15]. ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ( ) ( ( where: 27 (2.58) ) ) ) (2.59) √ √ ( ) ( ) ( ( ) ( ) [ ( ( )] ) ( ) ( ) ( ( 2.3 ) ) ) Bidirectional Converter For battery system applications, a bidirectional converter is required that can interface between the battery voltage and the DC-link voltage, which is typically of much greater magnitude. Unlike converters for fuel cell or PV application, in which current flows only from the source (low voltage) side to the DC link (high voltage) side, current in a battery system must be able to flow in either direction to charge or discharge the battery. The FIBC could be adapted for this purpose. If the four diodes (D1-D4) were replaced with additional switching devices with complementary switching signals, current would be able to flow in either direction. This is illustrated in Figure 2.8. The PSIM simulation results of the power flow, shown in Figure 2.9, illustrate the ability of this topology to go from charging (negative power flow) to discharging (positive power flow) at time t = 2s. 28 Figure 2.8 Bidirectional 4P FIBC. Figure 2.9 Power results for bidirectional 4P FIBC. 2.4 Conclusion This chapter presented the results of a quantitative simulation study of different, non-isolated DC- DC boost converter topologies. The results highlighted the advantages of the FIBC topology over other, conventional topologies. These included: improved efficiency, lower ripple, and reduced switching stress. Increasing the number of phases in the FIBC increases these benefits. However, the performance 29 advantages of the 6P over the 4P were marginal, and thus were deemed insufficient to justify the additional components required. For this reason, the 4P FIBC converter was chosen as the most promising candidate for this application. A method for developing a bidirectional version of this circuit was also presented with simulation results. A mathematical analysis of the 4P FIBC was conducted. The average state space equations for the six state variables (four inductor currents and two capacitor voltages) were developed for CCM operation. These were used to obtain the steady state, static transfer function for the output voltage. This verified that the 4P FIBC has a higher voltage gain compared to a conventional boost converter for a given duty cycle. Using the steady state, static transfer functions, the transfer functions for the inductor current vs. duty cycle and capacitor voltage vs. duty cycle were developed. These will be used to develop the linear feedback controllers in Chapter 4. In addition to regulating current and voltage, the controller must be able to execute MPPT. Chapter 3 provides an overview of different MPPT techniques, and a description of the chosen method for this prototype. 30 CHAPTER 3 MAXIMUM POWER POINT TRACKING Solar PV modules have a non-linear voltage-current relationship with a complex dependence of the PV module performance upon solar-irradiance intensity and PV module temperature. For part of their operational range, PV modules can be considered approximately current sources, while for the other portion of their range they can be considered to be voltage sources. The result of this is that there exists one point along the voltage/current curve for which the power produced by the PV module is at its maximum. Therefore, the MPP varies under varying irradiance and temperature conditions, and a realtime impedance matching with the load is necessary for optimization of the power conversion. Due to the relatively low conversion efficiency and high cost of PV technology, extracting the maximum useful power is an important design criterion for PV power systems. Therefore, a number of techniques have been developed in order to operate PV systems at their MPP. These techniques are known collectively as MPPT. 3.1 MPPT Algorithms Many different MPPT algorithms have been developed and used [20], [21]. Some are very simple, and can be easily accomplished using analog control techniques. One of the simplest is the CV technique, which maintains the PV array voltage at a predetermined limit, which corresponds to the MPP under a specific set of atmospheric conditions [21]. When operating outside of the predetermined conditions, this will only approximate the true MPP. Other approximate MPPT techniques include the short current pulse and open circuit voltage techniques. The SCP periodically shorts the PV module (or array). A current reference can be calculated as a percentage of the short circuit current (typically 92%) [21]. The OCV works similarly by open circuiting the array and determining a voltage reference based on the open circuit voltage (generally around 76% of the open circuit voltage) [21]. Each of these suffers the disadvantage of requiring additional switches to short or open the PV modules (or arrays). In addition, no power is produced during the period in which the panel is in a short or open condition [21]. Another approximate method, known as the beta method, uses an intermediate variable, . This variable is defined by the following equation [26]. ( ) where the constant c is related by: 31 (3.1) (3.2) Here, q is the electron charge, η is the diode ideality factor of the panel, KB is the Boltzmann constant, T is the rated temperature, and Ns is the number of PV cells in series. This method works by continuously calculating reference value using measurements of the voltage and current, and comparing it against a The MPPT algorithm can then operate using a conventional feedback control loop. One disadvantage to this method is that is a function of temperature. Therefore, this technique will not accurately track the MPP when operating outside rated temperature, unless c and are updated in response to temperature. Figure 3.1 P&O MPPT algorithm. These approximate MPPT methods are best suited for analog implementation. More accurate methods attempt to determine the exact MPP rather than an approximation, but are best implemented using a digital controller. One of these, the P&O method, works by inducing small changes in the operating voltage or current of the PV module and determining whether the power increases or decreases. If the induced change increases the power output, the operating conditions will be incremented in the 32 same direction. If the power decreases, the increments will be reversed. A flow diagram, shown in Figure 3.1, illustrates the decision structure utilized by the P&O technique. This method has a tendency to oscillate around the MPP, and can be ineffective under rapidly changing atmospheric conditions [20]. The IC method attempts to remedy these shortcomings [27]. It works based on the fact that the derivative of the power with respect to the voltage should zero at the MPP. This derivative can be expressed by [20]: ( ) (3.3) From this, it can be determined that: ⁄ at MPP ⁄ left of MPP ⁄ right of MPP (3.4) Using these inequalities, the MPP can be achieved. Additionally, the IC technique indicates both the direction and distance of the operating condition from the MPP. A modification to this technique uses a feedback control loop to drive ⁄ to zero [20]. The flow diagram in Figure 3.2 shows the decision structure for the IC technique [27]. The main disadvantage for the IC algorithm is that it is computationally complex, requiring many division operations. 33 Figure 3.2 IC algorithm flow diagram. Two exact MPPT techniques, P&O and IC, were implemented in simulation using both an analog and a digital controller. These were chosen because they have the best demonstrated performance [20], [21]. In addition, they do not require additional hardware components, such as switching devices to open or short the array, as are required by the open circuit voltage and short current pulse methods. These techniques are also not temperature dependent, which is the case for the beta technique [26]. The output of the MPPT was added to the current reference from the voltage controller (see Chapter 4). For the analog version, both methods used an increment/decrement interval (Δ) of 0.2 mA. For the digital implementation, the controller exhibited a greater tendency to overshoot. In order to compensate for this, Δ was scaled according to the output voltage error ( ( )). Thus, when the DC link voltage was far from the desired value, the MPPT used a larger Δ to converge more quickly. As the converter approached the nominal DC link voltage, Δ was decreased to minimize oscillations and deviations from the nominal voltage. This adaptation allowed the controller to quickly converge to the MPP while minimizing overshoot. However, it can only be applied in situations where no other device is regulating the DC link voltage. If another device is regulating this voltage (e.g., an inverter, battery, etc.) then this technique cannot be used. 34 One notable issue with implementing MPPT for the 4P FIBC was in accurately determining the total input current (IIN); many MPPT strategies require IIN in order to work correctly. For the FIBC, IIN is the sum of the inductor currents (where P is the number of phases) minus the load current: ∑( ) (3.5) This is because the load (output) current is common to both the inductors on the non-floating and floating phases. Therefore, in order to obtain the input current, a fifth current sensor must be added to measure either the total input current, or the output current, so that the input current can be obtained as in (3.5). As the magnitude of the output current would be significantly lower, it would be more economical to measure this current. 3.2 Conclusion MPPT is necessary to extract the maximum usable power from solar PV systems operating under varying temperature and irradiance conditions. Many different MPPT methods have been developed and implemented. These techniques vary from simple, approximate methods, which are best suited for analog implementation. More advanced, exact methods can be implemented using digital controls. The P&O method has a tendency to oscillate around the MPP, and can be negatively affected by rapidly changing atmospheric conditions; the IC method attempts to remedy these shortcomings. When implementing these MPPT, the total input current is required. In the case of the FIBC, an additional current sensor is required. Either the input or output current must be measured in addition to the four inductor currents to obtain this value. Chapter 5 presents the simulation results for the IC and P&O techniques under various operating conditions, using both analog and digital control algorithms. The development of the analog and digital controllers is discussed in Chapter 4. 35 CHAPTER 4 CONTROL DESIGN The controllers were designed for the 4P FIBC using a frequency analysis based on Bode diagrams. Proportional-integral (PI) controllers were used for the voltages and currents. The controller gains were defined by the crossover frequency (fc) and phase margin (PM), such that infinite gain was related to zero steady-state error, and the crossover frequency was related to the setting time of the compensated system. First, an analog controller was developed, using the Bode diagram frequency analysis. Then, a discrete controller was developed based on the analog controller. 4.1 Analog Control Design The analog controller had two outer voltage control loops that independently regulated the two output capacitor voltages in order to achieve the desired DC link voltage. This used the fact that the output DC link voltage of the FIBC is the series combination of the two capacitor voltage, minus the input (PV) voltage, or: (4.1) The reference value for the two capacitors can be expressed as a function of the desired output voltage (VDC*) and input voltage by: )⁄ ( (4.2) This approach is different from the dual loop controller used by Kabalo, et al., in [16], in which the outer loop controller regulated the total DC link voltage. Garcia, et. al., in [17] used a similar strategy in which both output capacitor voltages were controlled independently. For this application, controlling the two capacitors individually was found to yield better performance, particularly in low irradiance condition and the corresponding voltage-current response of the PV. The open-loop input and output voltage response for the 4P FIBC is shown in Figure 4.1. The figure illustrates that past a certain point, increasing the duty cycle causes the PV to enter its current source region, and the voltage to decrease. The decreased input voltage leads to decreased output voltage. 36 Figure 4.1 Open loop input and output voltage response. If only the output voltage is controlled, an overshoot in the controller will cause the voltage to cross the knee-point of the curve, and then decrease. The controller will attempt to compensate by increasing the duty cycle, which will cause the output to go even lower, until the controller eventually saturates. In contrast, by controlling the capacitor voltages, the reference is dependent on the magnitude of the input voltage, as shown in (4.2). Therefore, the input voltage can act as a feed-forward term, making the controller more robust and allowing faster crossover frequency. The benefits of this strategy are most apparent in low irradiance conditions, where the knee point of the voltage curve occurs at a lower value of input current. This is illustrated in Figure 4.2, in which the irradiance steps from 200 to 300 W/m2 at time t=1 s. Figure 4.2 Power response for VC and VDC voltage control strategies for a step change in irradiance from 200 to 300 W/m2 at time t=1 s. 37 The capacitor voltage (VC) control scheme showed a much more stable response. The VDC control scheme showed steady state stability problems, and failed to converge following the step irradiance change. The output from the voltage controller [Cv(s)] then fed four, independent, PI current controllers [Ci(s)] that regulated the four inductor currents. An additional term from the MPPT was added to this current reference. The outputs from the current control loops were then used as the modulation signals for four Pulse Width Modulation (PWM) controllers, which provided the gate pulse signals for the four MOSFET switches independently. The PWM controllers used 20 kHz switching frequency, where triangle wave carrier signals were phase shifted 90o apart. This control scheme topology is illustrated in Figure 4.3. Figure 4.3 Control block diagram for 4P FIBC. 38 4.2 Current Control Loop The previous studies supported the choice of a dual loop controller was used. A faster, inner current control loop integrated the MPPT functions, while an outer voltage control loop maintained the desired DC link voltage. The current and voltage control loops represented in Figure 4.3 can be incorporated in closed-loop, as indicated in Figure 4.4. All sensor gains, current (Ki) and voltage (Kv), were taken to be unity, as well as the PWM transfer function. Inner current control loop V*1,2 +- ev Cv(s) i*L +- ei Ci(s) Ki m iL PWM u Gv(s) v1,2 Gi(s) Kv Figure 4.4 4P FIBC control loops. The inner current control loop was designed first, considering only the closed control loop inside of the dashed box in Figure 4.4. The fc and PM were defined as 1.5 kHz and 75°, respectively. Using classic frequency control methods, the Bode diagram of the open loop response for the uncompensated and compensated system has been plotted in Figure 4.5. ( ) ( ) (4.3) 39 Figure 4.5 Bode diagram of the open loop uncompensated and compensated current control loop of the FIBC control scheme. 4.3 Voltage Control Loop After the current controller, the voltage control loop was designed, which depended on the inner current control loop. The desirable fc and PM were designed to be 5 Hz and 85°. This low crossover frequency was used to help with steady state stability; at higher values of fc, there were oscillations in the steady state current response. This may have been due to the voltage ripple across these capacitors; using low pass filters with a low fc might allow the bandwidth of this controller to be increased. The PI controller was designed based on these values for fc and PM, and the Bode diagram of the open loop, compensated system is shown in Figure 4.6, which showed that fc = 5 Hz and PM = 85°. ( ) ( ) (4.4) 40 Figure 4.6 Bode diagram of the open loop uncompensated and compensated voltage control loop of the FIBC control scheme. 4.4 Discrete Controller A discrete controller was developed using the previously developed analog controller. For the discrete controller, the analog PWM was replaced with a discrete PWM [25]. The PI controllers were replaced with discrete versions. Several methods for developing discrete versions of PI controllers exist [25]. For the backward Euler method output of the integration, ( ), utilizes the following expression: ( ) where is the integral gain, ( ) ( ) (4.5) is the sampling period, and ( ) is the error for sample k [25]. The digital expression of the complete PI loop can then be expressed as: ( ) where ( ) ( ) is the proportional gain. 41 ( ) ( ) (4.6) Another method of implementing a digital approximation of a PI controller uses a trapezoidal technique known as the Tustin method. This uses the weighted average of the present and previous error terms for the integration, reducing low frequency distortion [25]. This is expressed by: ( ) ( ) ( ) ( ) ( ) ( ) (4.7) Both the backward Euler and Tustin methods were employed in simulation and hardware. The Tustin appeared to perform better than the backward Euler technique, and was therefore used, although the performance difference was very slight. According to the Nyquist theorem, the sampling frequency should be more than twice (and up to an order of magnitude greater than) the switching frequency. However, sampling at this high rate would impose an unnecessarily high computational burden. Also, as the modulation index is only updated once per switching cycle, the feedback sampling rate should also be updated once per cycle. In order to obtain a working controller while only updating the state feedback values once per cycle, the signals can be filtered using a low pass filter to extract the average value. This filter must have a crossover frequency that is much lower than the switching frequency, which reduces the system’s bandwidth and increases the controller’s response time. Another solution is to synchronize the samples according to the switching pulses. Assuming that the ripple in the state variables is equal to the switching frequency, the average value occurs during the middle of the switch on or off period. By timing the sampling to occur during this time, the average value of the state variable can be obtained. This allows the controller to function without requiring low pass filtering to obtain the signal’s average value [25]. For the discrete controller, each inductor current was sampled at the middle of the switch on or off period. The voltages, which were subject to higher frequency ripple than the currents, had to be filtered through a low pass filter with a 50 kHz crossover frequency prior to sampling. By employing these strategies, all feedback variables could be sampled only once per switching cycle, thereby minimizing the computational burden. 4.5 Conclusion This chapter provided an overview of the development of a discrete, dual-loop, linear feedback controller for the 4P FIBC. First, an analog controller was developed, which was then discretized. Four independent, inner current loops regulated the current through the four inductors. An outer voltage loop controlled the voltage across the two output capacitors. Controlling the two capacitor voltages 42 independently was found to offer improved performance compared to controlling the complete DC link voltage. The output of the MPPT was added to the output of the voltage control loops. PI controllers were used for each control loop. The proportional and integral gain terms were calculated using frequency analysis based on Bode diagrams. The output of the four current control loops set the modulation index for the four PWM controllers. A discrete controller was developed from the analog controller. The analog controllers were replaced with discrete versions using the Tustin approximation method. By synchronizing the sampling times to occur at the center of the switch on or off pulse, the average value of the inductor currents could be obtained without the use of low-pass filters. A digital PWM replaced the analog PWM. Simulation results comparing the performance of the analog and discrete controllers, using two different MPPT techniques, are provided in Chapter 5. 43 CHAPTER 5 MODELING AND SIMULATION Computer based modeling and simulation techniques have been important to evaluate the controller performance. The control tools available in Matlab/Simulink (a block diagram based simulator) were used to assist in the control development, and final simulations were conducted using PSIM (a circuit based simulator). Both analog and discrete controllers were simulated to evaluate the performance of the discrete controller with respect to the analog version. Two different MPPT methods were simulated using both analog and discrete controllers. 5.1 Simulation Description Matlab/Simulink was used to develop the analog and digital controllers. After being designed in frequency domain, the analog controller was then implemented using PSIM software, and validated against the Simulink results. The discrete controller was simulated in C code using a C block function available in PSIM. Two MPPT methods, the P&O and IC techniques, were implemented using the analog controller. Their performance was evaluated in response to step changes in solar irradiance at various temperatures. The controllers were evaluated according to the total energy delivered to a resistive load, and the input current and voltage ripple. The output of each MPPT algorithm was added to the inductor current reference that came from the voltage control loops. A Δ of 0.2 milliamperes (mA) was used. Both MPPT algorithms were tested with two step changes in irradiance: 1) 1,000 W/m2 to 500 W/m2, and 2) 500 W/m2 to 600 W/m2. These two step changes were conducted at three different temperatures: 1) 10° C, 2) 25° C, and 3) 40° C. Thus, a total of six simulations were conducted for each algorithm. After conducting the simulations using the analog controller, the P&O and IC algorithms were also evaluated using the discrete controller, using the same six simulation scenarios. All the simulations were conducted using PSIM with a fixed simulation time step of 0.5 µs. The input to the 4P FIBC was a PV array based on 26 Kyocera KC200GT PV modules in a 2 series/13 parallel configuration. This PV module was selected because it has been well characterized and accurate models have been developed by Villalva, et al., in [28]. 44 The output, a resistive load, was varied in accordance with the power produced by the PV array, so that a constant voltage could be maintained under varying irradiance conditions. For the three irradiance conditions—1,000, 600, and 500 W/m2—the resistor values used were 27, 45 and 54 Ω, respectively. 5.2 Simulation Results The step change in the PV power output in response to a step decrease in λ from 1,000 W/m2 to 500 W/m2 for the P&O method using the analog linear feedback controller is plotted in Figure 5.1. The response for the IC was nearly identical, and has been omitted. Figure 5.1 Step response for analog P&O. When used with the analog controller, the P&O and IC techniques had virtually indistinguishable performance. Both delivered nearly identical amounts of energy to the load, and both were able to operate with very little ripple. Input voltage and current ripple were calculated to be less than 1% for both MPPT algorithms. Next, the P&O and IC algorithms were evaluated using the discrete controller. One issue that was encountered with the discrete controller implementation was that the voltage control loop tended to overshoot, driving the inductor current reference too high, and causing the system to saturate at the PV array’s short circuit current. This was particularly evident under low solar irradiance conditions. To compensate, the gain values for the voltage control loop were reduced. A reset for the integral term was introduced; when the PV voltage was detected to drop below a certain threshold (e.g., 20 VDC), the integral term of the voltage control loop was reset to zero. In order to increase the convergence time while avoiding overshoot, an adaptive MPPT method was developed. Rather than a fixed Δ, it was scaled according to the output voltage error. The Δ was set 45 to equal 5*10-6*(VDC*-VDC). In this way, the MPPT converged quickly but avoided overshoot. Figure 5.2 shows the PV power responses for a fixed versus a variable Δ using the discrete IC controller with a step change in irradiance from 1,000 W/m2 to 500 W/m2 at 25 oC. It shows that the variable delta was able to converge much more quickly following the change in irradiance. This technique, however, is only valid when only the PV boost converter is controlling the DC link voltage. If another device is regulating this voltage (e.g., an inverter or other DC converter) then this technique will no longer work, and a fixed Δ must be used. Figure 5.2 Fixed versus variable delta for IC MPPT, with step change in irradiance from 1,000 W/m2 to 500 W/m2 at t = 0.5 s. The PV power for the discrete P&O and IC MPPT controllers in response to a step change in irradiance from 1,000 W/m2 to 500 W/m2 at 25 oC is plotted in Figure 5.3 and Figure 5.4. In the discrete implementation, the P&O technique had a more observable ripple and tended to oscillate. The IC method, in contrast, was much smoother and more stable, particularly in response to the step change in irradiance at t = 1 s. 46 Figure 5.3 Step response for discrete P&O. Figure 5.4 Step response for discrete IC. The energy delivered to the load, in Joules (J), for the analog and discrete versions of the two MPPT methods is listed in Table 5.1. (Note: 1 J is 1 W/s.) While in the analog mode, input current ripple was very low (< 1%), the discrete P&O and IC methods had input current ripple values of 6.36% and 2.90%, respectively. Ripple is increased because of the discretely sampled feedback values; in the analog mode, the MPPT controller could respond instantaneously to variations in the operating conditions. In contrast, when using the discrete controller, feedback variables were sampled only once per switching period. This increased ripple resulted in lower energy capture, as the ripple causes the PV to deviate from its MPP. Overall, the discrete P&O delivered on average 11.84% less energy. The discrete IC had better performance, delivering on average 1.69% lower energy. 47 Table 5.1 MPPT Simulation Energy Results Comparison T=10o C T=25o C T=40o C Step Step Step Step Step Step decrease increase decrease increase decrease increase P&O 4,514 3,312 4,287 3,151 4,058 2,988 IC 4,514 3,312 4,287 3,151 4,056 2,988 P&O 3,718 2,762 3,789 2,940 3,694 2,703 IC 4,358 3,159 4,250 3,132 4,013 3,001 Technique Analog Discrete Percentage Difference P&O -17.63% -16.61% -11.62% -6.70% -8.97% -9.54% IC -3.46% -4.62% -0.86% -0.60% -1.06% +0.44% These results show that the P&O and IC methods had nearly identical performance when implemented with the analog controller. In the discrete version, both MPPT algorithms delivered less energy to the load, particularly at low temperature. However, the P&O performance was much more degraded with the discrete controller compared to the IC method. 5.3 Conclusion These simulation results validated the effective performance of the analog and discrete controllers for the 4P FIBC. They demonstrated that both the P&O and IC MPPT algorithms were able to effectively respond to a step change in irradiance. When using the analog controller, in which state feedback 48 variables were continuously updated, both MPPT algorithms had virtually indistinguishable responses. When the system was discretized, and feedback variables were discretely sampled once per switching cycle, the performance of both MPPT algorithms decreased, as measured by the amount of energy delivered to the load. However, the P&O method was more severely impacted by the discrete controller than the IC. Overall, the discrete P&O delivered nearly 12% less energy as compared to the analog version. The IC, in contrast, had a less than 2% reduction in energy capture compared to the analog version. Therefore, the IC method appears to be the better solution for implementation in a digital microcontroller. Chapter 6 describes the development of the hardware prototype and experimental setup. 49 CHAPTER 6 EXPERIMENTAL PROTOTYPE DEVELOPMENT A hardware prototype was constructed based on a Microchip dsPIC33FJ256GP710A embedded microcontroller. The prototype was designed for a low voltage input in the range of 48-72 VDC, with a nominal 400 VDC output at 5 kW. The implemented control functions were: sampling analog feedback signals, implementing the dual-loop linear feedback control algorithms, performing MPPT, and providing the switching signals to the MOSFETs. Four LEM LAH-25 NP Hall Effect Current Transducers (CTs) measured the inductor currents, while four LEM LV-20 P Hall Effect voltage sensors monitored the input, output and capacitor voltages. The outputs from the measurement devices were passed through op-amp signal conditioning circuits that scaled the outputs to match the analog measurement range of the microcontroller, and provided active low pass filtering. An overcurrent protection circuit sensed if any inductor current exceeded 50 A. If an overcurrent was detected, the switching pulses were disabled, and a 200 A relay was opened to isolate the fault. Overvoltage protection was also included. If the output voltage exceeded 475 V, a 2 kΩ, 100 W resistor was connected to the DC link by an IGBT to provide dynamic braking. Finally, four gate drivers had to be designed. These receive the switching signals from the microcontroller and provided the necessary voltage and current to switch the MOSFETs. Table 6.1 Selected Component Ratings and Part Numbers. Name Cin C1, C2 Description Rating Input 330 µF capacitor 350 V Output 1,000 µF capacitors 250 V Power D1-D4 L1-L4 Q1-Q4 diodes Inductors 50 A 250 µH 50 A MOSFET 650 V switches 84 A 50 Part number LNT2V331MSEC ECOS2EP102DX RHRG560 N/A STW88N65M5 Selected device ratings and part numbers are given for key components in Table 6.1. The inductor components were custom-manufactured, as commercially available inductors that met the desired current, inductance, and operational frequency specifications could not be found. Other hardware related activities included: thermal management and heat sink sizing, signal and power isolation between different parts of the circuit, and developing basic, proof of concept (POC) systems. 6.1 Inductor Custom inductor circuit elements had to be manufactured and procured, as no commercially available inductors were available which satisfied the following criteria: 20 kilohertz (kHz) operational frequency Average current of 34.4 Amperes (A) Peak current of 41.1 A Inductance value of 250 micro-Henreys (µH) at full rated current Maximum operating temperature of 180 oC Magnetics, Inc. provided eight (8) Kool Mu E cores, which are distributed air gap cores made from a ferrous iron alloy powder for low losses at elevated frequencies [29]. Magnetics, Inc. 00B802001 bobbins were also provided. These materials were delivered to Badger Magnetics for manufacturing. The Kool Mu 00K8044E026 core has a nominal inductance factor (AL ) of 91 +/-8% nano-Henrys per Telsa squared (nH/T2) . Using the worst case value of 83.72 nH/T2 and a desired inductance (L) value of 250 µH, the initial number of turns (N) was determined to be: √ √ (6.1) where L is specified in nano-Henrys (nH). Next, the DC bias in Ampere turns per centimeter (AT/cm) must be calculated, using: (6.2) 51 where is the path length of the core, in centimeters (cm). Based on the permeability vs. DC bias curve found in [29], the permeability at this bias would be roughly 75% of nominal, yielding an inductance value of roughly 250*0.75=187.5 µH at peak rated current. In order to compensate, the DC inductance value should be increased by increasing the number of turns. This is done iteratively by using different values for N, calculating the DC bias, and determining the derated inductance. Using this process, a total of 68 turns was determined to be required. This would yield an open-circuit inductance of 421 µH +/-8%, and at full rated current the inductance would be 295 µH +/8%. The test report, included in Appendix A, shows that the open circuit inductance of the four inductors was between 422.8-436.4 µH. The DC resistance (DCR) for all inductors was 22 mΩ. It was desired to maintain a peak current density of 500-600 A/cm. Using 9 AWG wire with a cross sectional area of 70.2*10-3 cm with a peak current of 41.1 A yields a peak current density of 585.5 A/cm, which is within the desired range. The core was able to accommodate 68 turns of 9 AWG wire with a 35.1% fill factor. Fill factor can be calculated by: (6.3) where is the cross-sectional area of the wire and is the window area of the core. This fits within the range for “Full winding,” which spans fill factors between 30-45% [29]. After winding, the completed coils were varnished with Class H insulation. The final inductors included 6” (15.24 cm) long, 8 AWG, Teflon wire leads that were terminated with soldered ring terminals. The four completed inductors are shown mounted to the chassis in Figure 6.1. The input capacitor, bus bars, and heat sink are also visible. 52 Figure 6.1 Completed inductors attached to chassis 6.2 Thermal Management Initially, IGBT semiconductor switching devices were chosen for the design because of their high power handling capacity and ability to withstand voltage transients. The IRG7PH35UD IGBT from International Rectifier (IR) was originally planned to be used, and was implemented as part of the first POC system. However, switching and conduction losses were determined to be too high with these devices. Conduction losses are approximately equal to the collector-emitter voltage during the conduction period (VCE(on)) times the average current through the device times the duty cycle. Through simulation, the average current under rated power and voltage conditions was determined to be 27.6 A, with a duty cycle of 0.75. The typical value for VCE(on) was given as 1.9 V on the manufacturer’s data sheet; the maximum value was 2.3 V. Using the worst case, conduction losses were calculated as: ( (6.4) ) 53 Switching losses for the IGBTs were calculated by the sum of the turn-on energy (EON) and the turn-off energy (EOFF) times fsw. The values for EON and EOFF came from the manufacturer’s data sheet, and 20 kHz was used for the value of fsw. From this, switching losses were calculated as: ( ) ( ) (6.5) Total power dissipation is the sum of conduction and switching losses, or 105 W per transistor. The basic equation for heat transfer can be written as: (6.6) where: TJ = the junction temperature in oC. (150 oC maximum) TA = ambient air temperature in oC. (30 oC used) RθJC = thermal resistance from junction to case of the semiconductor device in oC per Watt (oC/W) (0.7 oC/W, from manufacturer’s data sheet). RθCS = thermal resistance through the interface between the semiconductor device and the surface on which is it mounted in oC/W. (0.35 oC/W from silicon mounting pad data sheet). RθSA = thermal resistance of the heat sink in oC/W. From this equation, the required thermal resistance of the heat sink for a single IGBT can be specified by using: ( ) ( ) (6.7) This is an unrealistically small number. No commercially available heat sinks could be identified that were capable of meeting this requirement. Moreover, this is for a single device. If multiple devices 54 were to be mounted to a single heat sink, thereby increasing the power dissipation, the required thermal resistance of the heat sink would have to be a negative value. For this reason, the IR IRG7PH35UD IGBT was replaced with the STMicroelectronics STW88N65M5 MOSFET. This differed from the previous laboratory FIBC prototypes, which used IGBTs [5], [17]. Conduction losses for the MOSFET were calculated by: ( ( ) ) (6.8) Switching losses were calculated using: ( ) ( ) ( ) (6.9) where VDS is the drain to source voltage, ID is the drain current, tr and tf are the rise and fall times (from manufacturer’s data sheet), and tsw is the switching period (1/fsw). From these, total power dissipation for the STW88N65M5 was calculated to be 19.46 W, less than one fifth of the losses compared to the IGBT. Power dissipation in the Fairchild Semiconductor RHRG5060 power diode consisted only of the conduction loss, and was calculated by: ( ) (6.10) Two heat sinks were used, each supporting two MOSFETS and two power diodes. Total power dissipation was 67.92 W per heat sink. To maintain a maximum junction temperature of 150 oC, with an ambient air temperature of 30 oC, the required heat sink thermal resistance was determined to be: ( ( ) ) Here, RθJC was taken as the average of the MOSFET (0.28 oC/W) and the diode (1.0 oC/W). 55 (6.11) The 10.08” extruded aluminum heat sink from Heatsink USA, LLC has a thermal resistance of approximately 0.8 oC/W/3”. Doubling the length will decrease the thermal resistance by approximately 2/3. Therefore, a 6” piece of this heat sink would have a thermal resistance of approximately 0.53 oC/W. This provides a significant margin of safety, as it is about 32% below the minimum required thermal resistance. Changing the IGBT for a MOSFET necessitated a redesign of the gate driver. Because MOSFETs are more susceptible to transient overvoltages, the snubber was also redesigned, and a Transient Voltage Suppression (TVS) diode was added. 6.3 Proof-of-Concept Development Two POC systems were constructed and tested. These were low fidelity, laboratory test circuits used to evaluate and debug the power stage and ancillary circuits. The first POC system is shown in Figure 6.2. A single stage boost converter was implemented using an available inductor. This was not the custom-manufactured inductor that was used for the final prototype. This system was constructed prior to changing the IGBT for a MOSFET, so the IRG7P35UD and its associated gate driver were used. The input source was a DC power supply. The switching signal was provided by a pulse generator. A detailed look at the breadboard is provided in Figure 6.3. 56 Figure 6.2 POC 1. Figure 6.3 POC 1 annotated breadboard. After changing from the IGBT to the MOSFET devices, a second POC was constructed to evaluate the performance of the new semiconductor and gate driver. This did not include the sensor or signal conditioning circuitry, or the overcurrent protection. A 12 V car battery, protected by a 25 A fuse, served as the source. In this prototype, the switching signal came from the dsPIC33F microcontroller, rather than the pulse generator. The custom-manufactured inductors for the final prototype had been received from the manufacturer; one of these was used as the inductor for the second POC system. A 5 µF, high frequency, polypropylene (PPE) capacitor was added in parallel with the existing electrolytic output capacitor to minimize output voltage transients. Figure 6.4 shows the second POC system in operation. The microcontroller provided the pulses for the switching signal. Once again, this system was operated in open loop. The microcontroller sampled a 0-10 V analog voltage source, and set the duty cycle from 0-1 based on this external signal. The incoming 12 V from the battery is stepped up to 35.3 V, a voltage gain of about 3, at a duty cycle of 0.75. 57 Figure 6.4 POC 2. Figure 6.5 shows the breadboard for the second POC system. This much smaller board consisted of an: inverting and (nand) gate, optocoupler (for signal isolation), gate driver, and snubber, along with the diode and MOSFET. Figure 6.5 POC 2 annotated breadboard. 58 6.4 Snubber Development The purpose of the snubber is to minimize transients during switching operations. The snubber consisted of a capacitor and resistor connected in series across the transistors collector and emitter (for IGBTs) or drain and source (for MOSFETS). Because they are designed to damp high frequency transients, the capacitors and resistors used must be tolerant of voltage and current transients. For this reason, PPE film capacitors and carbon composite resistors were used. For IGBT, which was less sensitive to these transients, snubber values were chosen arbitrarily as 100 nF for the snubber capacitor (CS) and 1 kΩ, 5 W for the snubber resistor (RS). The waveforms, in Figure 6.6 show minimal transients. Figure 6.6 VCE (yellow) and IL (blue) waveforms for POC 1, duty cycle = 0.75. For this test, the input source was the 12 V car battery, and the Tranex Badger custom inductor was used. The switching source came from a BK Precision 3300 Pulse Generator. The input voltage and current were 11.8 V and 16.6 A, and the output was 36.1 V and 4.35 A, for an overall efficiency of 80.2% When the IGBT was replaced with a MOSFET in POC 2, problematic transients were observed, particularly, at higher duty cycles. 59 Figure 6.7 VDS (yellow) and IL (blue) waveforms for POC 2, duty cycle = 0.75. For this test, the input source was the 12 V car battery, and the Tranex-Badger custom inductor was used. The switching pulses came from the dsPIC33F microcontroller. The large transient observed in the inductor current waveform when the MOSFET was turned off caused the gate driver to fail and made obtaining accurate measurements difficult. The sinusoidal oscillation observed in the VDS waveform was the result of a longer 12 AWG wire connection between the diode and output capacitor, and not the result of the snubber. Input voltage and current were 11.84 V and 14.44 A and output voltage and current were 35.0 V and 4.21 A, for an overall efficiency of 86.2% Based on the quick snubber design technique described in [30], CS should be about twice the sum of the output capacitance of the MOSFET (COSS) and the estimated mounting capacitance. At the time, the IR IRFP4868 MOSFET was being considered, which has a COSS is about 612 pF according to the manufacturer’s data sheet. Mounting capacitance was estimated to be 40 pF. Therefore, CS should be around 1,304 pF. Thus, a 1,500 pF PPE capacitor was selected and procured. In the end, however, the ST Microelectronics (STM) STW88N65M5 MOSFET was selected because of its higher voltage tolerance, making it less susceptible to transient overvoltages. This MOSFET has a COSS is around 223 pF, which means that the 1,500 pF snubber capacitor was oversized. However, it was not deemed necessary to resize 60 this capacitor for the new MOSFET; the consequence of this being slightly oversized is the potential for increased power dissipation in the snubber. The snubber resistor size was determined based on the drain to source voltage across the transistor (VDS), which is about 230 V for an output voltage of 400 V and an input voltage of 60 V, and ID, which was determined to be around 32 A. Thus, RS was calculated to be: ⁄ ⁄ (6.12) The power dissipation required by the resistors can be calculated based on the stored energy in the capacitor, or: (6.13) Carbon composite resistors where chosen due to their improved transient performance compared to the more common carbon fiber. As 1.5 W or 2 W carbon composite resistors were not found to be readily available for the desired resistance, two 15 Ω, 1 W carbon composite resistors in parallel were used for the snubber resistors. Figure 6.8 shows the improved waveforms for POC 2 with the redesigned snubber. Note that for these waveforms, the resolution on the current probe was decreased from 100 mV/A to 10 mV/A in order to ensure that the entire current transient was captured and not clipped due to the output range of the current probe. This shows that the improved snubber was able to reduce the magnitude and severity of the current transients, alleviating the interference with the gate driver and measurement instruments. Reducing the length of the 12 AWG connecting the diode to the output capacitor eliminated the ringing in the voltage waveform. Input voltage and current were 11.9 V and 15.2 A and output voltage and current were 37.1 V and 4.47 A, for an overall efficiency of 91.7%. These results validated the correct performance of the gate driver and snubber circuits. Overall, POC 2, which used the MOSFET, had much better efficiency compared to POC 1, using the IGBT. This validates the decision to switch from the IGBT to the MOSFET. 61 Figure 6.8 VDS (yellow) and IL (blue) waveforms for POC 2, duty cycle = 0.75. 6.5 Gate Driver The gate driver circuit is shown in Figure 6.9. The input signal (u1) had to first pass through an inverting and (nand) gate (U1). The other input of the nand gate was an enable bit; if the enable bit went low, the switching inputs were disabled (see Section 6.7). The switching signal then passed through an RC circuit to a Fairchild HCPL2531 optocoupler (U2). This optocoupler allowed the signal ground to be decoupled from the gate driver, which was referenced to the source of the MOSFET. A 12 V source connected through a 5.6 kΩ pull-up resistor (R2) provided a 0-12 V input to the STM L6387E gate driver (U3). This is a dual channel gate driver, but in this instance only the low-side driver was used. The gate driver was connected to the MOSFET (Q1) through a gate resistor (R3). Like the snubber resistor, this was a carbon composite material in order to be more robust against voltage transients. An anti-parallel diode (D2) allowed the transistor to turn off more quickly. Two zener diodes (Z1 and Z2) prevented the gate to source voltage (VGS) from exceeding the specified maximum of +/-25 V, and protected the gate against transient overvoltages. 62 Figure 6.9 Gate driver circuit schematic. 6.6 Analog Signal Conditioning The analog signal conditioning circuits had two primary goals: 1) to provide a 0-3.3 V input that could be read by the microcontrollers analog input channels, and 2) to provide noise rejection. This was done using a three stage op-amp analog circuit. The CT circuit is shown in Figure 6.10. Figure 6.10 CT signal conditioning circuit schematic. The CTs were LEM LAH 25-NP closed loop, Hall effect sensors that provided a 0-55 mA output for currents from 0-55 A (1 mA/A). This output current passed through a 100 Ω resistor (R1) to ground, thus generating a 0-5.5 V signal (100 mV/A). This passed through a TL074 quad op-amp (U1). The first stage of the op-amp was a non-inverting, unity gain buffer to provide high input impedance. The second stage was an inverting op-amp with a 10 kΩ trimming potentiometer (trimpot) (P1) that allowed for variable gain. The final stage was a unity gain, inverting op-amp to negate the inversion of the previous stage. It contained a capacitor in the feedback loop in order to act as an active low pass filter, with fc near 200 kHz, or 10 times the switching frequency (20 kHz), as given by: 63 (6.14) The voltage sensor signal conditioning circuit is similar, and is shown in Figure 6.11. Figure 6.11 Voltage sensor signal conditioning circuit schematic. A LEM LV-20P Hall Effect voltage sensor has been used, which worked based on a 0-14 mA input. An appropriately sized resistor was therefore provided to produce the necessary input current based on the magnitude of the voltage to be measured. The Hall Effect voltage sensor produced a 0-35 mA output that was passed across a 133 Ω resistor to provide a 0-4.655 V output. Again, this passed through a three stage op-amp circuit, using a TL074 quad op-amp. Because the voltages are subject to higher ripple than the inductor currents, a 3,300 pF capacitor was used with a 1 kΩ resistor to provide a low pass filter with a 48.2 kHz fc. 6.7 Overcurrent Protection The overcurrent protection circuit used four comparator circuits to compare the voltage output from the four CTs against a reference value. If the current in any one of the inductors exceeded the reference, it caused the normally high (5 V) output to go low. This then caused the output of the logic inverter (U4) to go high. U4 is connected to the clock input of a D-type, positive edge triggered flip-flop (U5), causing it to change its output Q from high (5 V) to low. As this was an edge triggered device, it acted very quickly. The output Q was an enable signal that was connected through the nand gate in the gate drive circuit (see Section 6.5). The inverse output of U5 ( ̅ ) sent a disable signal to the microcontroller, and opened Relay 2. Relay 2 breaks the neutral connection for the 12 V supply that powered the coil for the 200 A input relay. The emergency stop button (SW2) is also located on this neutral path. The circuit returns to normal by pushing the reset button (SW1). 64 Figure 6.12 Overcurrent protection circuit schematic. 65 6.8 Overvoltage Protection The overvoltage protection system worked on the principle of dynamic braking. A rotary potentiometer (P1) and a 1 kΩ resistor formed a voltage divider to scale VDC. A Schmitt trigger comparator circuit (U1) compared this against a 5 V reference value. P1 changed the midpoint of the hysteresis and the trimpot controlled the high and low trigger values. These were set to turn on at 475 V and turn off at 425 V. When the overvoltage circuit turns on, transistor Q1 provided a 12 V on signal to the IR2117 gate driver circuit (U3). U3 is used to drive the IR IRG7PH35UD IGBT. The IGBT was used for this circuit because of its higher voltage rating, and the fact that IGBTs are more robust against voltage transients. Q1 closed the neutral connection on a 2 kΩ, 100 W resistor, R10. The additional output impedance helped to drive the DC link voltage back towards its nominal value. Figure 6.13 Overvoltage protection circuit schematic. 66 6.9 Complete Circuit The complete circuit consists of: Power stage Four gate drivers Four analog current signals Four analog voltage signals Overcurrent protection Overvoltage protection This is shown in Figure 6.14. The complete components list is provided in Appendix B. 67 Figure 6.14 Complete circuit schematic diagram. 68 69 6.10 Hardware Design The FIBC prototype was built using a 16.75” x 23.5” (42.5 x 59.7 cm) rack for a rack mount enclosure. The incoming low voltage DC connected to two 1” (2.54 cm) wide, 1/4” (.635 cm) thick, copper bus bars. The positive bus bar was broken by an OMRON G9EC-1 DC12 200 A, normally open relay. The 12 V source providing the coil voltage for this relay could be broken in the event of an overcurrent, or by pushing the emergency stop button. Each bus bar had lugged connections to two inductors; L1 and L2 were connected to the positive bus bar, L3 and L4 were connected to the negative bus bar. Additionally, two 8 AWG copper wires connected the neutral bus bar to the Printed Circuit Board (PCB), where they connected to the source of transistors Q1 and Q2. Two more 8 AWG wires connected the positive bus bar to the drain of Q3 and Q4. A pair of 12 AWG wires from the line (positive) side of the 200 A relay connected to a terminal block on the PCB to provide voltage to the isolated power supplies, so that the on-board controls could be activated prior to the application of primary power. A 330 µF, 350 V, chassis mount input capacitor was installed between the positive and neutral bus bars on the load (negative) side of the 200 A relay. A 40 kΩ resistor in parallel with the input capacitor to provide a path for inductor currents to flow if Relay 2 was opened. All primary power connections to the PCB were soldered, 12 AWG wire. The 8 AWG wire from the inductors and bus bars were terminated on a terminal block. From the terminal block, each 8 AWG was transmitted by 2 12 AWG wires to the PCB in order to facilitate soldering. The PCB measured 10” x 16” (25.4 x 40.64 cm) and was placed between two 10.08” (25.6 cm) heat sinks. The heat sinks were affixed the rack with angled aluminum brackets. The dsPIC33F microcontroller was used with the Microchip Explorer 16 board. This board was placed in a steel enclosure that was installed above the PCB, supported by additional angled aluminum brackets. An additional piece of aluminum was installed across the two heat sinks at the front of the rack to hold the emergency stop switch, the reset switch for the overcurrent protection, the chassis mount PPE capacitors, and the panel-mount potentiometer for the overvoltage protection circuit. The high voltage output was accessed from the front of the PCB; two two-circuit, 60 A Molex terminal blocks were connected to the positive and negative outputs. A two layer PCB was designed, using 2 ounce per square foot (oz/ft2) copper thickness and 1/8” FR4 dielectric. The thicker dielectric was chosen to minimize electromagnetic interference and thermal heat transfer between layers. This thickness, however, impeded soldering. For future systems, it is 70 recommended to use a thinner dielectric. The board was organized with high current carrying traces along the outer edges, and control and signal processing circuits near the interior. The high current traces were required to handle a peak current of 41.1 A. For external traces in air with 2 oz/ft 2 copper weight, the trace width must be 511 mils in order to limit the temperature rise of these traces to 30 oC [31]. The PCB is shown in Figure 6.15. Figure 6.15 PCB for 4P FIBC. The microcontroller was on the Explorer 16 development board from Microchip. It was connected to the PCB by 15-pin D-sub connectors. The 15 pins were connected to: Eight analog inputs (four voltages, four currents) One digital input (disable) Four digital outputs (four PWM switching pulses) Two common wires. A picture of the hardware prototype is shown in Figure 6.16. 71 Figure 6.16. Hardware prototype. 6.11 Isolation and Electromagnetic Interference One technical challenge that was unique to this circuit was the multiple numbers of electrical neutral (common) points within the circuit. A total of five common points existed within the circuit; these are defined as: Input source neutral (GND 0) – this will most likely be earth-grounded in most installations Signal ground for control electronics (GND 1) – this was kept isolated with respect to the input ground to prevent circulating currents within the control electronics DC link negative (GND 2) – this was floating with respect to the input The source of MOSFET Q3 (GND 3) – the source of this transistor was floating below ground potential. The associated gate driver must be referenced to this to provide the proper VGS to turn the transistor on and off. The source of MOSFET Q4 (GND 4) – see previous. In addition, the chassis was connected to earth ground, which was kept isolated from all the other commons. To maintain isolation, a number of isolated power supplies were required to provide control voltage with respect to the different common potentials. Small, isolated, switching DC-DC power 72 supplies were specified and procured which had a wide input range, from 18 – 75 VDC, and could therefore be powered from the input source. This would have allowed the hardware prototype to be completely self-powered. The PCB with the switching DC-DC supplies is shown in Figure 6.17. Figure 6.17 PCB with switching DC-DC converters. MOSFETs Q1 and Q2 had their sources at the same potential as the input, so a 12 V power supply referenced to this common was required for their associated gate drivers. The control electronics required +/- 12 V and 5 V, so supplies referenced to the signal ground were provided. A switching DCDC supply provided +/- 12 V, and a linear 5 V regulator provided the 5 V using the +12 V supply. The overvoltage protection circuitry for the DC link had to be with respect to the DC link negative, and required 12 V and 5 V; the 12 V was provided by a switching DC-DC supply, and the 5 V was from a linear 5 V regulator. Finally, the gate drivers for Q3 and Q4 each needed their own isolated 12 V power supply referenced to their sources. Optocouplers were used to maintain isolation between the gate signals originating from the microcontroller, which was referenced to the signal ground, and the gate drivers, which were all referenced to the source of their associated MOSFETs. Although the switching DC-DC supplies were convenient, allowed the prototype to be completely self-powered, and required very little space on the PCB, they were found to produce significant Electromagnetic Interference (EMI). The EMI was most noticeable in the analog feedback signals from the CTs and Halle Effect voltage sensors, particularly in the 350 kHz and 2 Megahertz (MHz) range. As 73 no other circuits were operating at this frequency, the switching power supplies were identified as the likely culprit. The magnitude of these transients was so severe that it caused the overcurrent protection to act, disabling the system. In order to resolve this issue, the switching DC-DC supplies had to be removed, and replaced with external linear power supplies running off of line voltage. The prototype system, with linear power supplies, is shown in Figure 6.18. Figure 6.18 Hardware prototype with external linear power supplies. Replacing the switching DC-DC supplies with linear power supplies did resolve the EMI issue. However, this change led to some notable disadvantage: 1) the prototype was no longer self-powered – 120 VAC line voltage was now required, 2) the prototype was now bulkier due to the linear power supplies, 3) set-up and tear-down of the prototype became more complicated and time consuming due to the external power connections. Therefore, in order to mitigate these disadvantages, as well as the EMI issue, future systems should install the switching DC-DC supplies on a separate PCB housed inside a shielded enclosure. This will allow the system to enjoy the advantages of the switching DC-DC supplies, notably, being smaller, lighter, and completely self-powered. Housing the supplies separately in a shielded enclosure will prevent the EMI from impacting the control and/or protection systems of the FIBC. 74 In addition to the EMI produced by the switching DC-DC supplies, EMI was generated by the switching transients for each MOSFET, most noticeably in the current and voltage measurement feedback signals. To prevent these noise transients from impacting the microcontroller, it was run separately on its own board from the main PCB, and housed inside a steel electrical enclosure. For the low power conditions under which testing was conducted, this EMI did not appear to adversely impact system performance. At high power, however, these transients could begin to interfere with the analog feedback and/or digital switching pulses to a degree that may be detrimental. Further reductions in EMI could be accomplished by: Improved snubbers to reduce switching transients Changing the connector between the microcontroller and the PCB from 15-pin to 26-pin so that each signal can be a twisted pair with its own common Separating the control circuitry from the power circuitry and shielding 6.12 Microcontroller Code The microcontroller code for the dsPIC33FJ256GP710A was written primarily in C, with some assembly code, using Microchip MPLAB IDE v8.70 and compiled for the device using the MPLAB C30 C compiler. The dsPIC was run at its maximum operating speed of 40 Mega-Instructions Per Second (MIPS), using an external 8 MHz oscillator and the on-chip Phase-Locked-Loop (PLL). The four, phase-shifted PWM outputs were generated using a timer with a 2,048 cycle period to generate a 19.53 kHz switching frequency. This lower frequency was used (as opposed to exactly 20 kHz) because 2,048 is a power of two, which had computational time benefits when using the fixed-point calculations (more on this later). In order to stagger the four PWM outputs, the Continuous Pulse Mode of the Output Compare channels had to be used, rather than the normal PWM mode for the microcontroller. Typically, when using digital PWM, the output is set high at the start of each counter period and goes low when the counter value matches the modulation index. However, for this particular dsPIC, only two of the nine timers were available for PWM. For this reason, there was no way to stagger the pulses 90o apart using the normal PWM mode. The Continuous Pulse mode allowed the high and low times for each channel to be set independently. Although this mode is typically associated with fixed, not variable, pulse widths, it did not appear to have any adverse impacts on operation. Additional code was required to implement this switching mode. For future work, it is recommended to investigate the Motor Control (MC) class of dsPIC33F microcontrollers to see if they can accommodate more flexibility in PWM configuration. 75 Four additional Output Compare channels were configured with periods half the length of the PWM channels. These triggered an Interrupt Service Routine (ISR) at the center of the on or off period of each switching pulse. For low duty cycle (< 40%), the ISR was triggered at the center of the off pulse, and for high duty cycle (>40%) is occurred at the center of the on pulse. This allowed the sample to be taken on whichever slope was smaller, thus minimizing skewing error. Inside this ISR, the corresponding inductor current was sampled, the PI control calculations for that current were executed, and the duty cycle for the next switching period was written. The entire computation time for each current samplecontrol ISR was approximately 5.6 microseconds (µs). An additional ISR was triggered using a timer with a period of ten switching cycles (512 µs). This ISR sampled the input voltage and two capacitor voltages and ran the voltage control loops. This interrupt had a lower priority than the current sample-control ISRs, so that the microcontroller would continue to execute the current control ISRs at the correct times. This used the Nested Interrupt capability of the dsPIC microcontroller. The microcontroller required approximately 38.8 µs to sample the three analog inputs and run the control equations. For these experiments, which were conducted using a fixed resistive load, it was not possible to have the MPPT and voltage control loops working simultaneously. A fixed output voltage with a fixed resistive load would result in fixed power, in which case the MPPT could not function. Therefore, the MPPT was included in a separate program. First the CV method was implemented, which maintained the input voltage at a fixed value. The CV method is one of the simplest MPPT methods to perform, and was used as a baseline against which other methods could be evaluated. The input voltage control loop used the same techniques and gain terms as the output voltage capacitor control. The MPPT control loop was in a separate ISR that was executed every 2,048 µs. In addition to the CV algorithm, the IC method was also implemented. The output of the IC MPPT was a voltage reference for the input voltage controller used by the CV algorithm, using a variation (Δ) of 41.7 millivolts (mV). The total computation time required to run the IC algorithm was about 39 µs. As was mentioned in Chapter 3, the MPPT required the total input current in order to work correctly. This was approximated as the sum of the inductor currents minus the output voltage divided by the load resistance. All calculations were executed using fixed-point arithmetic to allow for very fast computation time, using 16 bit signed integers. The controller must be modified as follows in order to function with fixed point integers. 76 Figure 6.19 Fixed-Point Integer Control Block Diagram The controller gain values were determined for unity gain for the feedback signals, and a modulation index varying between zero and one. In reality, however, the analog feedback signals were subjected to scaling as they were converted to 10 bit integers by the Analog to Digital Converter (ADC). All mathematical calculations were performed using 16 bit signed integers; therefore, the modulation index varied from zero to 215 (32,768). Integral anti-windup was inherently present because after the controller exceeds this value it would reset to zero. In order to adapt the original gain terms for the fixed point integer controller, they had to be scaled by m/ Ki where m is the maximum value of the modulation index (215) and Ki is the signal gain caused by the ADC. To adapt this modulation index to the PWM carrier varying from zero to a peak value COUNTP, it had to be scaled by COUNTP /m. Making this value a power of two was beneficial for computational speed; for this reason, the counter period was increased from 2,000 (corresponding to a switching frequency of 20 kHz) to 2,048 (211) (corresponding to a switching frequency of 19.53 kHz). Finally, the scaled modulation index was written to the duty cycle register for the PWM; the duty cycle was limited to 85% for safety and protection reasons. 6.13 Cost The complete bill of materials for the hardware prototype is included as Appendix C. The total cost for all parts, components, and materials was approximately $2,367. High volume pricing quotes for all major components were obtained from manufacturers and/or vendors. It is estimated that with volume manufacturing, the material cost could be reduced to about $1,330 (roughly 44%). Assuming that the material cost represents 60% of the total manufacturing cost, the total cost for commercial production of this device is estimated to be about $2,217. 6.14 Conclusion A hardware prototype of the 4P FIBC was constructed. Custom inductor components that were capable of meeting the circuit specifications had to be designed and manufactured. Thermal dissipation calculations led to the eventual use of MOSFET devices over IGBTs, due to the lower losses within the MOSFETs. These calculations were used to appropriately size and select a heat sink. Two POC systems 77 were constructed and tested to validate certain aspects of the design and debug issues with the circuits, including the snubber. A snubber was designed to minimize transients that occurred during the switching operations of the MOSFETs. In addition to the power stage, ancillary circuits had to be designed and built. These included four isolated gate drivers capable of receiving the switching pulses from the microcontroller and driving the MOSFETs. Analog signal conditioning circuits were required to scale and filter the CT and voltage sensor measurement signals and provide them to the microcontroller. Overcurrent and overvoltage protection were implemented as well. The chassis for the prototype was a 16.75” x 23.5” (42.5 x 59.7 cm) rack for a rack mount enclosure. The low voltage DC was connected to a 1” x 1/4” (2.54 x 0.635 cm) bus bar. The positive bus bar was broken by a 200 A, normally open relay contact. The overcurrent protection would trip this relay to isolate the source in the event of a fault. The PCB was installed between two 10.08” (25.6 cm) heat sinks that were affixed to the chassis. The dsPIC33 microcontroller was run on the Microchip Explorer 16 board, which was placed above the PCB in an aluminum enclosure. The microcontroller was responsible for executing all control functions for the prototype. These included: sampling analog feedback signals, implementing the dualloop linear feedback control algorithms, performing MPPT, and providing the switching signals to the MOSFETs. Fixed point integer calculations were used to reduce computational time. The total cost for the prototype was $2,357. For full commercial production, the total estimated system cost is approximately $2,210, including manufacturing. A few design issues were identified that should be noted, and can easily be rectified in future designs. First, a thinner dielectric (e.g., 1/16”) should be used to facilitate ease of soldering. Second, EMI from the switch-mode DC-DC supplies must be mitigated by placing these devices on a separate board and using metal shielding. Third, additional EMI reductions could be accomplished by improved snubbers for the MOSFETs and/or separating and isolating the control circuitry from the power stage. Finally, changing the connector between the microcontroller and the PCB from 15-pins to 26-pins would allow each signal to be paired with its own common, allowing them all to be twisted pairs, further reducing EMI. Chapter 7 describes the experimental testing and results using the prototype described in this chapter, where a comprehensive evaluation and comparison of simulated and experimental data is performed. 78 CHAPTER 7 EXPERIMENTAL TESTING AND RESULTS Experimental testing of the hardware prototype that was described in Chapter 6 was conducted to verify correct operation and validate the control algorithms that were described in Chapter 5. First, open loop operation was conducted to verify the correct operation of the gating signals, gate drivers, and power stage. Figure 7.1 shows the four PWM pulses from the microcontroller, which were staggered 90 o (12.8 µs) apart. Next, the steady state and step response of the current controller was demonstrated. After that, the outer voltage control loop was implemented. Finally, the MPPT was applied and validated. All testing was conducted under low power conditions. This was due to the fact that no low voltage, high current input source was available, nor was there a suitable high power load bank capable of withstanding 400 VDC. Figure 7.1 Staggered PWM pulses. 7.1 Current Controller The current controllers were tested using a 900 W HP 6439B DC power supply for the input and a 33 Ω resistive load bank at the output. The experimental test setup for these tests is shown in Figure 7.2. Each of the four stages of the converter was tested individually, then all four together. Figure 7.3 shows the four inductor currents under steady state conditions for a 3 A reference. The signals came from the onboard CTs, which provided an output of 100 mV/A. Note that EMI from the switching transients can be 79 observed in these output signals. Figure 7.4 shows two inductor currents (IL1 and IL3) along with the input current (IIN). Figure 7.2 Experimental test setup for current control testing. Figure 7.3 Four inductor currents, steady state, 3 A reference. 80 Figure 7.4 Inductor currents IL1 and IL3 and input current IIN, steady state, 3 A reference. Note that although the inductor currents have the characteristic triangle wave shape typical of a conventional boost converter, the effect of the interleaving was to yield an input current that was much more linear. Input current ripple was calculated to be approximately 8.6% under these operating conditions, whereas the current through the inductors had about 67% ripple. The step response for a single inductor current, IL2, for a step change in reference from 4 to 8 A at time t = 2 ms is plotted in Figure 7.5. The plot shows the current waveform as measured by a current probe, plotted at 1 A per division. Also plotted are the external current reference signal (I*), provided by a function generator, and the error signal from the microcontroller (e I2). The time is plotted at 2 ms per division. As the microcontroller did not have a digital to analog converter, the error signal was written as the duty cycle for an unused PWM channel, and filtered through an RC low pass filter to obtain the average value. The scaling factor for this signal was 477 mV/Amp error, and it was plotted at 200 mV/division. The step response for the change in reference from 8 A down to 4 A is plotted in Figure 7.6. Figure 7.7 shows several cycles of the inductor current responding to the step change in reference. 81 Figure 7.5 Step response for inductor current IL2 for a reference from 4 to 8 A at time t = 2 ms, 1 A/division, 2 ms/division. Figure 7.6 Step response for inductor current IL2 for a reference from 8 to 4 A at time t = 2 ms, 1 A/division, 2 ms/division. 82 Figure 7.7 Step response for inductor current IL2 for a reference from 4 to 8 A, multiple cycles. The step response for all four inductor currents is shown in Figure 7.8. Because the oscilloscope only had four channels, the reference is not plotted. Figure 7.9 shows the same condition with the reference plotted in place of IL3. Figure 7.10 shows multiple cycles of the four inductor currents responding to the step change in reference value. Figure 7.8 Step response for four inductor currents for a reference from 2.5 to 4 A at time t = 1 ms, 2 A/division, 1 ms/division. 83 Figure 7.9 Step response for four inductor currents (IL3 not shown) for a reference from 2.5 to 4 A at time t = 2 ms, 2 A/division, 2 ms/division. Figure 7.10 Step response for four inductor currents for a reference from 2.5 to 4 A, multiple cycles, 2A/division, 100 ms/division. These results show that the inner, current control loop was able to quickly and accurately respond to a step change in input with minimal overshoot. 84 7.2 Voltage Controller The voltage controller was tested under low voltage, low power conditions due to limitations on input current and load dissipation capability. The input DC source provided 20 V; the 4P FIBC prototype increased this to 140 V at the output (a voltage gain of 7). The resistive load consisted of a 157 Ω load bank and a 300 Ω resistor connected in parallel by a switch. This switch allowed the output load to be varied from 157 to 103 Ω in order to validate the ability of the voltage controller to maintain output voltage under a varying load. The experimental setup for the voltage control testing is shown in Figure 7.11. Figure 7.11 Experimental test setup for voltage control testing. The output voltage response is plotted in Figure 7.12. These measurements were taken using the output of the Hall Effect voltage sensor, after the analog signal conditioning. The scale factor for this is 169.7 V/V. The waveform was plotted on the oscilloscope at 50 mV/division (which corresponds to 8.5 V/division when considering the scale factor), 50 ms/division. At the moment when the output load was decreased from 157 to 103 Ω, the voltage sagged to 131.3 V (6.2%); when the load returned to 157 Ω the voltage rose to 15.4 V (8.9%). The transient for the load decrease was approximately 76 ms, while when the load resumed the transient lasted approximately 136 ms. 85 Figure 7.12 Output voltage response to step change in load from 157 Ω to 103 Ω and back, 50 mV/division, 50 ms/division. Figure 7.13 shows the result for the input current, measured by a current probe, and three of the four inductor currents (IL1, IL2, and IL4, as measured by the on-board current sensors. (Note that these sensors produce an output voltage of 100 mV/A.) Figure 7.14 shows the four inductor currents, as measured by the Hall Effect current sensors. Figure 7.13 Input and inductor current response for a step change in load from 157 Ω to 103 Ω and back, 2 A/division, 100 ms/division. 86 Figure 7.14 Four inductor current response for a step change in load from 157 Ω to 103 Ω and back, 2 A/division, 100 ms/division. These results show that the voltage controller was able to effectively regulate the output voltage of the 4P FIBC under a varying load. The controller overshoot was acceptable—within 10%. While the control response was somewhat slow, it was deemed adequate for solar PV applications, which do not experience extremely rapid changes in operating conditions. Using a low pass filter to eliminate the ripple from the capacitor voltage feedback signals could allow the bandwidth of the voltage controller to be increased, which would improve performance. 7.3 Maximum Power Point Tracking Finally, the MPPT control capability was evaluated. As a PV input source was unavailable, this was tested by using Thevinin’s theorem for maximum power transfer, which states that for a source with a series impedance, the maximum power that can be transferred to a series load is achieved when the load impedance is equal to the source impedance. The 4P FIBC reduces the load resistance as seen by the input according to: ( ( 87 ) ) (7.1) By placing a resistor in series with the DC input power supply, the MPPT was expected to vary the duty cycle such that the load resistance as seen by the input (RIN) was equal to the series source resistance. The experimental test setup with the source resistance is shown in Figure 7.15. The input power supply was set for 40 V. Two 15 Ω, 4.47 A rheostats, connected in parallel, were used for the source resistance. A 102 Ω resistive load was used for the output. Figure 7.15. MPPT testing setup. First, the CV algorithm was tested. CV is not the most effective algorithm when using a PV input source, as the maximum power voltage varies with operating temperature. For this testing scenario, however, CV would always yield the correct answer if it was set to maintain the input voltage to the converter at half the source voltage. This made the CV method very useful for the purposes of debugging, calibrating, and evaluating the IC method. The source resistance was varied between 2-4 Ω. Table 7.1 compares the input power for the CV and IC MPPT algorithms compared to the theoretical value. The CV results were taken to be correct, with any discrepancy attributed to imprecision in the rheostat resistance, and the IC results were evaluated against these. These results show that both algorithms were able to converge to operating points at or near the actual maximum power for each value of source resistance. The IC algorithm results were very close to the CV results for source resistance of 3 Ω and 4 Ω, but diverged for 2 Ω. This is because the IC algorithm allowed the input voltage to rise for low source impedance, causing it to deviate from the maximum power point. It is anticipated that for a PV input source, which has higher input impedance, this will perform better. This is supported by simulation results. 88 Table 7.1 MPPT power comparison. Source Input Power Resistance (Calculated) CV Input Power IC Input Power Percent Difference 4Ω 100 W 99.4 W 98.1 W 1.3% 3Ω 133.3 W 132.8 W 132.3 W 0.4% 2Ω 200 W 203.8 W 182.9 W 10.3% Figure 7.16 shows the simulation results for input power for a duty cycle sweep when a mathematical model of a PV array was used as the input source, and the input power result using the IC MPPT method. This figure shows that, when using the PV model as an input source, the IC MPPT converged to the maximum input power. However, when this array was replaced with a DC source with a series resistor, the results shown in Figure 7.17 indicate that the same IC MPPT was unable to converge to the maximum input power, but instead converged to a slightly lower value. For this reason, although the IC method failed to converge to the maximum power point in these experimental results, it believed that this algorithm will be capable of reaching the maximum power point when operating with a PV input source. Figure 7.16 Simulation result for IC MPPT method with duty cycle sweep using PV model input source. 89 Figure 7.17 Simulation result for IC MPPT method with duty cycle sweep using DC input source with series resistance. Figure 7.18 shows the input voltage and current response for the CV and IC algorithms in response to a change in source resistance from 2 Ω to 3 Ω. The input voltage was measured using the on board Hall Effect voltage sensor. This was scaled using the 26.303 V/V scale factor; a five point moving average was used to reduce the appearance of the ripple resulting from switching frequency EMI that was present in the sensor output. The input current was measured using a current probe. Note that the CV algorithm holds the input voltage steady at 20 V (half the source voltage) and only the current varies. This matches the expected performance for this algorithm, and translates to the maximum power under these particular operating conditions. In contrast, for the IC algorithm, both input voltage and frequency vary in response to the changing source resistance. This resulted in lower power capture for the IC method as compared to the CV, as plotted in Figure 7.19. 90 Figure 7.18 Input voltage and current response for CV and IC MPPT algorithms in response to a change in source resistance from 2 Ω to 3 Ω. 200 Power [W] 180 160 140 120 Power (CV) Power (IC) 100 0 2 4 6 8 Time [s] Figure 7.19 Power response for CV and IC MPPT algorithms in response to a change in source resistance from 2 Ω to 3 Ω. The input voltage and current response for the CV and IC MPPT algorithms is plotted in Figure 7.20. Once again, the CV was able to hold the input voltage constant at 20 V, while for the IC method this voltage varied with source resistance. The input power response for the two methods is plotted in Figure 7.21. 91 Figure 7.20 Input voltage and current response for CV and IC MPPT algorithms in response to a change in source resistance from 4 Ω to 3 Ω. Figure 7.21 Power response for CV and IC MPPT algorithms in response to a change in source resistance from 4 Ω to 3 Ω. These results show that both the CV and IC methods were able to converge at or near the maximum power conditions for varying values of source resistance. The CV method was useful for providing a baseline against which the IC results could be compared. The dynamic results illustrate the ability of these two methods to respond to changing operating conditions and converge to the new maximum power. The IC method diverged from the optimal conditions; it is believed that this divergence is due in part to the different voltage/current relationship for a DC voltage source with a series resistor when compared to a PV source. Future testing with either a PV simulator or PV array will be necessary to evaluate the MPPT algorithms under more realistic testing conditions. 92 7.4 Conclusion These experiments successfully demonstrated all control algorithms for the 4P FIBC under low power conditions. High power testing was precluded due to the lack of a suitable low voltage, high current, high power input supply and high power, 400 VDC load bank. The four current controllers were able to respond very quickly to step changes. Comparing the input current against the individual inductor currents demonstrated that the interleaving of the FIBC was able to significantly reduce the input current ripple to less than 10%. The voltage controller was able to maintain the desired value with acceptable deviation during step changes in load. The voltage controller performance could be increased by using a low pass filter with a lower fc to more effectively remove the ripple from the voltage feedback signal, which would then allow the controller bandwidth and gain terms to be increased. Two MPPT algorithms, the CV and IC, were successfully demonstrated. These were both able to converge at or near the theoretical maximum power point when tested using a DC input source with series source resistance. Future testing efforts are necessary to better evaluate the prototype under high power conditions, and with a real or simulated PV input source. 93 CHAPTER 8 CONCLUSION AND FUTURE WORK This thesis presented an analysis, design, and implementation of a prototype 4P FIBC for PV power system applications. This converter was chosen as a promising candidate for PV because of its high voltage gain, improved efficiency, and reduced input ripple when compared against other nonisolated DC-DC boost converter topologies. These advantages have led the FIBC topology to be demonstrated for fuel cell and electric vehicle applications. These same advantages also make the FIBC an improved solution for PV power system applications. A simulation study of eight different, non-isolated DC-DC converts was conducted. Of these, the FIBC showed the best performance in terms of efficiency, input and output ripple, and switching stress. A mathematical analysis was also performed to derive the state space equations, steady state, static transfer function, and transfer function. These were used for developing the dual loop, linear feedback controller. A bidirectional FIBC was also presented. In addition to the voltage and current control, this converter also required MPPT capability for use with PV power systems. MPPT is required due to the non-linear current and voltage relationship of PV modules. For each set of operating conditions, there is one unique point at which the maximum power of the PV can be obtained. This point varies with irradiance and temperature, and the MPPT controller is responsible for dynamically tracking to obtain the optimal output. Two MPPT algorithms, the IC and the P&O, were compared in simulation, and the IC was implemented for the prototype. For the FIBC, the total input current is the sum of the four inductor currents minus the output (load) current. Therefore, for some MPPT methods that require the input current, another current measurement of either the input or output current must be made. The dual loop, linear feedback controller was designed using Bode diagram frequency analysis. Four faster, inner current control loops regulated the current through each inductor. Two independent voltage control loops, which were slower than the current controllers, regulated the voltage across the two output capacitors. Controlling the capacitors individually, as opposed to attempting to regulate the total DC link output voltage, was found to lead to more stable operation, particularly for a PV input experiencing low irradiance conditions. After developing the analog controllers, they were discretized in order to be implemented using an embedded microcontroller. The digital PWM used a single counter with four Output Compare channels operating in the Continuous Pulse method to generate four, independent PWM outputs that were phase shifted 90o apart from one another. The PI controllers were replaced with discrete versions using the Tustin integration approximation. 94 Modeling and simulation studies of the complete prototype, with current and voltage controllers and MPPT algorithms, were conducted. These explored the performance variation between two MPPT methods, the P&O and IC techniques, using both the analog and discrete controllers. With analog controls, the P&O and IC techniques had virtually indistinguishable performance. However, when using the discrete controller, the IC method was observed to converge more quickly and accurately, and without the steady state oscillation that occurred with the P&O technique. For this reason, the IC method was chosen for implementation with the hardware prototype. A hardware prototype of the 4P FIBC was constructed, using a Microchip dsPIC33FJ256GP710A embedded microcontroller. This prototype used MOSFETs as opposed to IGBTs, due to the loss calculations and estimated thermal dissipation for the IBGTs. Two small POC systems, which were conventional boost converters, were constructed and tested in order to better refine the final prototype and evaluate system performance parameters. This was particularly helpful in designing the snubber for the MOSFET devices. In addition to the power stage, gate drivers, analog signal condition, overcurrent protection, and overvoltage protection circuits were also designed. All of these circuits were implemented on a single 10”x16” (25.4 x 40.64 cm), two layer PCB with 2 oz/ft 2 copper weight and 1/8” FR4 dielectric. The inductors were mounted directly to the chassis, and the microcontroller was housed separately and run from the Microchip Explorer 16 Development Board. The final prototype was built using a 16.75” x 23.5” (42.5 x 59.7 cm) rack for a rack mount enclosure. This rack served as the chassis and held all the buswork, inductors, printed circuit board, heat sink, terminal blocks, etc. Mounting brackets allowed the Microchip Explorer 16 Development Board, which held the embedded microcontroller, to be placed above the main PCB, inside a steel electrical enclosure. One unique challenge for the 4P FIBC was the multiple common points within the circuit. The high voltage DC link output of the device is floating with respect to the input source. In addition, the MOSFETs for the floating phases of the converter have their source floating with respect to the output and to each other. In all, five common points existed within the circuit: the input neutral, the output neutral, the source of MOSFET Q3, the source of MOSFET Q4, and the control/signal common, which was kept isolated from all others to prevent circulating currents. In addition, the chassis was connected to earth ground, which was kept isolated from all the other commons. A number of isolated power supplies were required to provide the necessary control voltages with respect to the different common points. Several issues related to EMI were identified with this prototype. Originally, small, on-board DCDC power supplies were included to provide the various isolated control voltages that were required. 95 However, these were observed to be a major source of EMI. This EMI was so severe that noise transients triggered the overcurrent protection circuit, disabling the prototype. In order to resolve this, these had to be replaced with external, linear power supplies. However, the advantages of the DC-DC supplies are their smaller size, weight, cost, and the fact that no external AC power supply is required. For these reasons, it is recommended that future prototypes use the isolated DC-DC supplies, but house them separately and use adequate shielding to prevent EMI from interfering with the control or protection circuitry. In addition, EMI was generated by switching transients from the MOSFETs. These transients were especially noticeable in the outputs of the on board current and voltage sensors. Under the low power conditions at which the hardware testing took place, these did not appear to noticeably impact system performance. However, at higher power, these transients could begin to significantly interfere with either the analog feedback signals or the digital switching pulses. Future work should be focused on minimizing these transients. Soft-switching of the MOSFETs would be beneficial. If necessary, the control circuitry could be separated from the power stage entirely, and run on a separate, shielded PCB. The microcontroller code was written primarily in C, with some assembly language. All calculations were performed using 16-bit signed, fixed point integers. Fixed point integer arithmetic was beneficial for computation time. The sample/control loops for each inductor current took approximately 5.6 µs, and were run once per cycle. Sampling was synchronized to occur at the center of either the on or off switching pulse to extract the average value. The voltage sample and control loop was executed in a separate ISR that occurred every ten switching cycles (512 µs) and took about 38.8 µs; the MPPT algorithm was executed in another ISR with a 20 cycle (1,024 µs) period and required 39 µs. The nested interrupt capability of the dsPIC microcontroller ensured that the slower, outer controls (voltage and MPPT) did not interfere with the inner current control loops. The total cost for all parts, components, and materials for the prototype was approximately $2,357. It is estimated that with volume manufacturing, the material cost could be reduced to about $1,326 (roughly 44%). Assuming that the material cost represents 60% of the total manufacturing cost, the total cost for commercial production of this device is estimated to be about $2,210. Experimental testing of the hardware prototype was conducted to verify correct operation of all circuits and control functions. The current, voltage, and MPPT controls were tested all evaluated. The current controllers were demonstrated under steady state and step response conditions. Comparing the input current against the individual inductor currents demonstrated that the interleaving of the FIBC was able to significantly reduce the input current ripple to below 10%. The voltage control was evaluated with 96 a step change in load. It was able to maintain the desired output voltage with acceptable overshoot and settling time. Finally, two different MPPT algorithms, the CV and IC were demonstrated. These were both able to converge at or near the maximum power point when tested using a DC input source with series source resistance. Future experimental work should focus on evaluating the prototype under rated voltage and power conditions. EMI issues will have to be closely monitored, and mitigations may be required if EMI begins to degrade system performance with increased voltage and current. The performance of the MPPT algorithms must be evaluated using a real or simulated PV input. The FIBC must be demonstrated as part of a complete PV power system. This will involve using the 4P FIBC as an initial stage to raise the PV array voltage up to the requisite value for an inverter, and using the high voltage DC link output of the 4P FIBC as an input for an inverter. Finally, the development of a bidirectional hardware prototype, suitable for battery system applications, is necessary. 97 REFERENCES [1] K. Bolcar, and K. Ardani, “International Energy Agency Co-operative Programme on Photovoltaic Power Systems: Task 1 Exchange and dissemination of information on PV power systems,” National Survey Report of PV Power Applications in the Unites States, May 2011. [2] L. Sherwood, “U.S. Solar Market Trends 2012” Interstate Renewable Energy Council (IREC), Jul. 2013. [3] M. Bazilian, et al., “Re-considering the Economics of Photovoltaic Power,” Blomberg NEF: New York, USA (2012). [4] M. Kabalo, “Conception et realization de convertisseur DC-DC pour une chaine de traction electrique et/ou hybride,” Ph.D. thesis, Dept. Eng. Sci. and Microtechnology, Univ. of Technology of Belfort-Montbeliard, Belfort, France, 2012. [5] M. Kabalo, et al., “Experimental validation of high-voltage ratio low-input-current-ripple converters for hybrid fuel cell supercapacitor systems,” IEEE Trans. Veh. Techno., vol. 61, no. 8, Oct. 2012. [6] D. Coutellier, et al., “Experimental verification of floating-output interleaved-input DC-DC highgain transformer-less converter topologies,” Proc. IEEE PESC, 2008, pp. 562-568. [7] S. Choi, et al., “Analysis, design and experimental results of a floating-output interleaved input boost-derived DC-DC high-gain transformer-less converter,” IET Power Electron., vol. 4, iss. 1, pp. 168-180, Jan. 2011. [8] G. Brinkman, et al., “Grid modeling for the SunShot Vision Study,” NREL, Golden, CO, Rep. TP6A20-53310, 2012, 38 pp. [9] National Renewable Energy Laboratory, “Renewable electricity futures study,” NREL, Golden, CO, Rep. NREL/TP-6A20-52409, 2012, vol. 1-4. [10] P. Denholm, et al., “Bright future: solar power as a major contributor to the U.S. grid,” IEEE Power & Energy Magazine, vol. 11, no. 2, pp. 23-32, Mar./Apr. 2013. [11] R. Messenger and J. Ventre, Photovoltaic Systems Engineering, 2nd ed. Boca Raton, FL: CRC Press, 2004. [12] R. Teodorescu, M. Liserre, and P. Rodriquez, Grid Converters for Photovoltaic and Wind Power Systems, 1st ed. West Sussex, United Kingdom: John Wiley & Sons, 2011. [13] M.H. Taghvaee, et al., “A current and future study on non-isolated DC-DC converters for photovoltaic applications,” Renewable and Sustainable Energy Reviews, vol. 17, pp. 216-227, Oct. 2012. [14] N. Mohan, T.M. Underland, and W.P. Robbins, Power Electronics: Converters, Applications, and Design, 3rd ed., Wiley, 2003. [15] M. Kabalo, et al., “Experimental evaluation of four-phase floating interleaved boost converter design and control for fuel cell applications,” IET Power Electronics, vol. 6, no. 2, pp. 215-226, Feb. 2013. 98 [16] M. Kabalo et al., “Advanced hybrid dual loop control for multi-phases interleaved floating DC-DC converter for fuel cell applications,” Industry Applications Society Annual Meeting (IAS), 2012 IEEE, pp. 1–8, 2012. [17] F. Garcia, et al., “Modeling and Control Design of the Interleaved Double Dual Boost Converter,” IEEE Trans. Ind. Electron. vol. 60, no. 8, Aug. 2013. [18] H. Sira-Ramírez and R. Silva-Ortigoza, Control Design Techniques in Power Electronics Devices, London, England, Springer, 2006. [19] Z. Liang, “A High-Efficiency PV Module-Integrated DC/DC Converter for PV Energy Harvest in FREEDM Systems,” IEEE Trans. Power Electron., vol. 26, no. 3, Mar. 2011. [20] T. Esram, and P. L. Chapman, "Comparison of Photovoltaic Array Maximum Power Point Tracking Techniques," IEEE Trans. Energy Conversion, vol. 22, no. 2, pp. 439-448, Jun. 2007. [21] R. Faranda and S. Leva, "Energy comparison of MPPT techniques for PV Systems," WSEAS Transactions on Power Systems, vol. 3, no. 6, pp. 446-455 Jun. 2008. [22] R. Leyva, et al., “MPPT of Photovoltaic Systems using Extremum-Seeking Control”, IEEE Trans. on Aerospace and Electronic Systems, vol. 42, no. 1, Jan. 2006. [23] M. G. Simoes and N. N. Franceschetti, “Fuzzy optimization based control of a solar array system,” IEE Proc.-Electr. Power Appl., vol. 146, no. 5, pp. 552-558, Sep. 1999. [24] M. Simoes, et. al, "LCL Filter Design and Performance Analysis for Grid Interconnected Systems," IEEE Trans. Ind. Appl., vol. PP, no. 99, pp.1-7. [25] S. Buso and P. Mattavelli, Digital Control in Power Electronics, Morgan & Claypool, 2006. [26] M. de Brito, et al., “Evaluation of the Main MPPT Techniques for Photovoltaic Applications,” IEEE Trans. Ind. Electron., vol. 60, no. 3, pp. 1156-1167, Mar. 2013. [27] K.H. Hussein, et al., “Maximum photovoltaic power tracking: an algorithm for rapidly changing atmospheric conditions,” IEE Proc.-Gener. Trasm. Distrib., vol. 142., no. 1, pp 59-64, Jan. 1995. [28] M.G. Villalva, et al., “Comprehensive Approach to Modeling and Simulation of Photovoltaic Arrays,” IEEE Trans. Power Electron., vol. 24, no. 5, pp. 1198-1208, May 2009. [29] Magnetics, Inc. “Magnetics, Inc. Powder Core Catalog,” 2012. Online, Available http://www.maginc.com/File%20Library/Product%20Literature/Powder%20Core%20Literature/MagenticsPowder CoreCatalog2012.pdf. [30] R. Severns, “Design of Snubbers for Power Circuits,” 2006. Online, Available: Online, Available http://www.cde.com/tech/design.pdf, Accessed Jan 20, 2014. [31] B. Suppanz, “Trace Width Website Calculator,” 2007. http://www.4pcb.com/trace-width-calculator.html, Accessed Mar 17, 2014. 99 APPENDIX A INDUCTOR TEST REPORT 100 101 APPENDIX B COMPONENT LIST 102 Table B-1 Component List. Subcircuit Item Value/ Type C1 1,000 µF C2 1,000 µF C3 330 µF C4 1,500 pF C5 1,500 pF C6 1,500 pF C7 1,500 pF C76 5 µF C77 5 µF CT1 CT CT2 I or P Rating V Rating Manufacturer Manuracturer Part # PANASONIC ECO-S2EP102DX PANASONIC ECO-S2EP102DX NICHICON LNT2V331MSEC CORNELL DUBILIER 715P15256JD3 CORNELL DUBILIER 715P15256JD3 CORNELL DUBILIER 715P15256JD3 CORNELL DUBILIER KEMET 715P15256JD3 C4BTJBX4500ZAJ J C4BTJBX4500ZAJ J 25 ADC LEM USA LAH 25-NP CT 25 ADC LEM USA LAH 25-NP CT3 CT 25 ADC LEM USA LAH 25-NP CT4 CT 25 ADC LEM USA LAH 25-NP D1 Power diode 50 A FAIRCHILD RHRG5060 D2 Power diode 50 A FAIRCHILD RHRG5060 D3 Power diode 50 A FAIRCHILD RHRG5060 D4 Power diode FAIRCHILD RHRG5060 TRANEX BADGER N/A TRANEX BADGER N/A TRANEX BADGER N/A TRANEX BADGER STMICROELECTRONI CS STMICROELECTRONI CS STMICROELECTRONI CS STMICROELECTRONI CS N/A Main 250 VDC 250 VDC 350 VDC 600 VDC 600 VDC 600 VDC 600 VDC 700 VDC 700 VDC KEMET L1 250 µH L2 250 µH L3 250 µH L4 250 µH 50 A 41.1 ADC 41.1 ADC 41.1 ADC 41.1 ADC Q1 MOSFET 84 ADC Q2 MOSFET 84 ADC Q3 MOSFET 84 ADC Q4 MOSFET 84 ADC R1 15 Ω 1W OHMITE OA150KE R2 15 Ω 1W OHMITE OA150KE R3 15 Ω 1W OHMITE OA150KE R4 15 Ω 1W OHMITE OA150KE 100 VDC 100 VDC 100 VDC 100 VDC 650 VDC 650 VDC 650 VDC 650 VDC 103 STW88N65M5 STW88N65M5 STW88N65M5 STW88N65M5 Subcircuit Item Value/ Type I or P Rating R5 15 Ω R6 15 Ω R7 V Rating Manufacturer Manuracturer Part # 1W OHMITE OA150KE 1W OHMITE OA150KE 15 Ω 1W OHMITE OA150KE R8 R87 15 Ω 40 kΩ 1W 6.5 W OHMITE VISHAY OA150KE CW00540K00JE73 RELA Y1 -- 200 ADC 400 VDC OMRON G9EC-1 DC12 50 A 300 V MOLEX 38280-0112 MICROCHIP dsPIC33FJ256GP71 0A LITTELFUSE P6KE350A LITTELFUSE P6KE350A LITTELFUSE P6KE350A LITTELFUSE P6KE350A VISHAY K103K10X7RF5U H5 TB1 TERMINAL BLOCK U1 Microcontroller Z1 TVS 600 W Z2 TVS 600 W Z3 TVS 600 W Z4 TVS 600 W 350 VDC 350 VDC 350 VDC 350 VDC Gate Driver – 1 of 4 C8 10 nF C9 D5 0.47 µF 1N914 0.2 A 50 VDC 50 VDC 100 V D6 STTH102 1A 200 V R9 390 Ω R10 R11 NICHICON FAIRCHILD STMICROELECTRONI CS UPW1HR47MDD 1N914BTR 1/4 W STACKPOLE CF14JT390R 5.6 kΩ 1/4 W STACKPOLE CF14JT5K60 4.7 Ω 1/2 W STACKPOLE RC12JB4R70 FAIRCHILD STMICROELECTRONI CS MICRO COMMERCIAL CO MICRO COMMERCIAL CO HCPL2531 U2 Optocoupler 8 mA U3 Gate Driver 400 mA Z5 18 V Zener 18 VDC Z6 18 V Zener 18 VDC 0-30 VDC 17 VDC STTH102 L6387E 1N5355B-TP 1N5355B-TP Gate Driver – 2 of 4 50 VDC 50 VDC VISHAY K103K10X7RF5U H5 NICHICON UPW1HR47MDD FAIRCHILD STMICROELECTRONI CS 1N914BTR 1/4 W STACKPOLE CF14JT390R 1/4 W STACKPOLE CF14JT5K60 C10 10 nF C11 0.47 µF D7 1N914 0.2 A 100 V D8 STTH102 1A 200 V R9 390 Ω R10 5.6 kΩ 104 STTH102 Subcircuit Item Value/ Type I or P Rating R11 4.7 Ω 1/2 W U4 Optocoupler 8 mA U5 Gate Driver 400 mA Z7 18 V Zener 18 VDC Z8 18 V Zener 18 VDC V Rating 0-30 VDC 17 VDC Manufacturer Manuracturer Part # STACKPOLE RC12JB4R70 FAIRCHILD STMICROELECTRONI CS MICRO COMMERCIAL CO MICRO COMMERCIAL CO HCPL2531 L6387E 1N5355B-TP 1N5355B-TP Gate Driver – 3 of 4 50 VDC 50 VDC VISHAY K103K10X7RF5U H5 NICHICON UPW1HR47MDD FAIRCHILD STMICROELECTRONI CS 1N914BTR 1/4 W STACKPOLE CF14JT390R 5.6 kΩ 1/4 W STACKPOLE CF14JT5K60 4.7 Ω 1/2 W STACKPOLE RC12JB4R70 FAIRCHILD STMICROELECTRONI CS MICRO COMMERCIAL CO MICRO COMMERCIAL CO HCPL2531 C12 10 nF C13 0.47 µF D9 1N914 0.2 A 100 V D10 STTH102 1A 200 V R9 390 Ω R10 R11 U6 Optocoupler 8 mA U7 Gate Driver 400 mA Z9 18 V Zener 18 VDC Z10 18 V Zener 18 VDC 0-30 VDC 17 VDC STTH102 L6387E 1N5355B-TP 1N5355B-TP Gate Driver – 4 of 4 50 VDC 50 VDC VISHAY K103K10X7RF5U H5 NICHICON UPW1HR47MDD FAIRCHILD STMICROELECTRONI CS 1N914BTR 1/4 W STACKPOLE CF14JT390R 5.6 kΩ 1/4 W STACKPOLE CF14JT5K60 4.7 Ω 1/2 W STACKPOLE RC12JB4R70 FAIRCHILD STMICROELECTRONI CS MICRO COMMERCIAL CO MICRO COMMERCIAL CO HCPL2531 C14 10 nF C15 0.47 µF D11 1N914 0.2 A 100 V D12 STTH102 1A 200 V R9 390 Ω R10 R11 U8 Optocoupler 8 mA U9 Gate Driver 400 mA Z11 18 V Zener 18 VDC Z12 18 V Zener 18 VDC 0-30 VDC 17 VDC CT Signal Conditioning – 1 of 4 105 STTH102 L6387E 1N5355B-TP 1N5355B-TP Subcircuit I or P Rating Item Value/ Type C16 82 pF C17 100 nF C18 100 nF P1 10 kΩ 1/2 W R21 10 kΩ 1/4 W R22 10 kΩ 1/4 W R23 4.7 kΩ 1/4 W R24 100 Ω 1/2 W U10 QUAD OP AMP V Rating 50 VDC 50 VDC 50 VDC Manufacturer VISHAY Manuracturer Part # K820J15C0GF5TH 5 K104K10X7RF5U H5 K104K10X7RF5U H5 BOURNES INC. 3362P-1-103LF STACKPOLE TEXAS INSTRUMENTS CF12JT100R VISHAY VISHAY TL074CN CT Signal Conditioning – 2 of 4 50 VDC 50 VDC 50 VDC C19 82 pF C20 100 nF C21 100 nF P2 10 kΩ 1/2 W R25 10 kΩ 1/4 W R26 10 kΩ 1/4 W R27 4.7 kΩ 1/4 W R28 100 Ω 1/2 W U11 QUAD OP AMP VISHAY K820J15C0GF5TH 5 K104K10X7RF5U H5 K104K10X7RF5U H5 BOURNES INC. 3362P-1-103LF STACKPOLE TEXAS INSTRUMENTS CF12JT100R VISHAY VISHAY TL074CN CT Signal Conditioning – 3 of 4 50 VDC 50 VDC 50 VDC C22 82 pF C23 100 nF C24 100 nF P3 10 kΩ 1/2 W R29 10 kΩ 1/4 W R30 10 kΩ 1/4 W R31 4.7 kΩ 1/4 W R32 100 Ω 1/2 W U12 QUAD OP AMP VISHAY K820J15C0GF5TH 5 K104K10X7RF5U H5 K104K10X7RF5U H5 BOURNES INC. 3362P-1-103LF STACKPOLE TEXAS INSTRUMENTS CF12JT100R VISHAY VISHAY TL074CN CT Signal Conditioning – 4 of 4 C25 82 pF C26 100 nF 50 VDC 50 VDC C27 100 nF 50 106 VISHAY K820J15C0GF5TH 5 K104K10X7RF5U H5 VISHAY K104K10X7RF5U VISHAY Subcircuit Item Value/ Type I or P Rating P4 10 kΩ 1/2 W R33 10 kΩ 1/4 W R34 10 kΩ 1/4 W R35 4.7 kΩ 1/4 W R36 100 Ω 1/2 W U13 QUAD OP AMP V Rating VDC Manufacturer Manuracturer Part # H5 BOURNES INC. 3362P-1-103LF STACKPOLE TEXAS INSTRUMENTS CF12JT100R TL074CN PT Signal Conditioning – 1 of 4 C28 3,300 pF C29 100 nF C30 100 nF C31 100 nF P5 10 kΩ 50 VDC 50 VDC 50 VDC 50 VDC 1/2 W 500 VDC PT1 VISHAY K332K15X7RF5TL 2 K104K10X7RF5U H5 K104K10X7RF5U H5 K104K10X7RF5U H5 BOURNES INC. 3362P-1-103LF LEM USA LV 20-P MFR-25FBF52133R VISHAY VISHAY VISHAY R37 1 kΩ 1/4 W R38 1 kΩ 1/4 W R39 4.7 kΩ 1/4 W R40 133 Ω +/-1% 1/4 W YAEGO R41 6.2 kΩ 2W U14 QUAD OP AMP STACKPOLE TEXAS INSTRUMENTS RSF2JT6K20 TL074CN PT Signal Conditioning – 2 of 4 C32 100 nF C33 100 nF C34 100 nF C35 3,300 pF P6 10 kΩ 50 VDC 50 VDC 50 VDC 50 VDC 1/2 W 500 VDC PT2 VISHAY K104K10X7RF5U H5 K104K10X7RF5U H5 K104K10X7RF5U H5 K332K15X7RF5TL 2 BOURNES INC. 3362P-1-103LF LEM USA LV 20-P VISHAY VISHAY VISHAY R42 40 kΩ 6.5 W VISHAY R43 133 Ω +/-1% 1/4 W YAEGO CW00540K00JE73 MFR-25FBF52133R R44 4.7 kΩ 1/4 W R45 1 kΩ 1/4 W R46 1 kΩ 1/4 W U15 QUAD OP AMP TEXAS TL074CN 107 Subcircuit Item Value/ Type I or P Rating V Rating Manufacturer INSTRUMENTS Manuracturer Part # PT Signal Conditioning – 3 of 4 C36 100 nF C37 100 nF C38 100 nF C39 3,300 pF P7 10 kΩ 50 VDC 50 VDC 50 VDC 50 VDC 1/2 W 500 VDC PT3 VISHAY K104K10X7RF5U H5 K104K10X7RF5U H5 K104K10X7RF5U H5 K332K15X7RF5TL 2 BOURNES INC. 3362P-1-103LF LEM USA LV 20-P VISHAY VISHAY VISHAY R47 22 kΩ 5W TE CONNECTIVITY R48 133 Ω +/-1% 1/4 W YAEGO 1623782-7 MFR-25FBF52133R R49 4.7 kΩ 1/4 W R50 1 kΩ 1/4 W R51 1 kΩ 1/4 W U16 QUAD OP AMP TEXAS INSTRUMENTS TL074CN PT Signal Conditioning – 4 of 4 C40 100 nF C41 100 nF C42 100 nF C43 3,300 pF P8 10 kΩ 50 VDC 50 VDC 50 VDC 50 VDC 1/2 W 500 VDC PT4 VISHAY K104K10X7RF5U H5 K104K10X7RF5U H5 K104K10X7RF5U H5 K332K15X7RF5TL 2 BOURNES INC. 3362P-1-103LF LEM USA LV 20-P VISHAY VISHAY VISHAY R52 22 kΩ 5W TE CONNECTIVITY R53 133 Ω +/-1% 1/4 W YAEGO 1623782-7 MFR-25FBF52133R R54 4.7 kΩ 1/4 W R55 1 kΩ 1/4 W R56 1 kΩ 1/4 W U17 QUAD OP AMP TEXAS INSTRUMENTS TL074CN Overcurrent Protection Comp 1 C44 10 nF C45 10 nF C46 10 µF 50 VDC 50 VDC 50 VDC 108 KEMET C315C103M5U5C A C315C103M5U5C A PANASONIC EEU-FC1H100 KEMET Subcircuit Comp 2 Comp 3 Comp 4 Main Item Value/ Type I or P Rating P9 10 kΩ R57 22 Ω R58 10 kΩ 1/4 W R59 10 kΩ 1/4 W C47 10 nF C48 10 nF P10 10 kΩ 1/2 W R60 10 kΩ 1/4 W R61 10 kΩ 1/4 W C49 10 nF C50 10 nF C51 10 µF P11 10 kΩ R62 V Rating Manufacturer Manuracturer Part # 1/2 W BOURNES INC. 3362P-1-103LF 1/4 W YAEGO CFR-25JB-52-22R 50 VDC 50 VDC 50 VDC 50 VDC 50 VDC KEMET C315C103M5U5C A C315C103M5U5C A BOURNES INC. 3362P-1-103LF KEMET KEMET C315C103M5U5C A C315C103M5U5C A PANASONIC EEU-FC1H100 1/2 W BOURNES INC. 3362P-1-103LF 22 Ω 1/4 W YAEGO CFR-25JB-52-22R R63 10 kΩ 1/4 W R64 10 kΩ 1/4 W C52 10 nF C53 10 nF P12 10 kΩ 1/2 W R65 10 kΩ 1/4 W R66 22 Ω 1/4 W C54 10 µF C55 100 nF C56 1 µF C57 100 nF C58 100 nF C59 100 nF D13 LED 13 mA R67 22 Ω 1/4 W R68 4.7 kΩ 1/4 W R69 10 kΩ 1/4 W R70 390 Ω 1/4 W R71 560 Ω 1/4 W 50 VDC 50 VDC 50 VDC 50 VDC 50 VDC 50 VDC 50 VDC 50 VDC 5 VDC 109 KEMET KEMET C315C103M5U5C A C315C103M5U5C A BOURNES INC. 3362P-1-103LF PANASONIC VISHAY EEU-FC1H100 K104K10X7RF5U H5 K105K20X7RF5U H5 K104K10X7RF5U H5 K104K10X7RF5U H5 K104K10X7RF5U H5 KINGBRIGHT WP710A10ID5V YAEGO CFR-25JB-52-22R KEMET VISHAY VISHAY VISHAY VISHAY Subcircuit Item Value/ Type I or P Rating R72 RELA Y2 1 kΩ 1/4 W SW1 SW2 U18 U19 U20 U21 U22 U23 U24 2 ADC PUSH BUTTON SWITCH EMERGENCY STOP DUAL COMPARATOR DUAL COMPARATOR DUAL COMPARATOR HEX LOGIC INVERTER 2A 10 ADC V Rating Manufacturer Manuracturer Part # 30 VDC PANASONIC DS2E-SL2-DC5V JUDCO 40-2531-01 OMRON A22E-M-01 FAIRCHILD LM393N FAIRCHILD LM393N FAIRCHILD TEXAS INSTRUMENTS TEXAS INSTRUMENTS TEXAS INSTRUMENTS LM393N TOSHIBA ULN2803 PANASONIC VISHAY EEU-FC1H100 K104K10X7RF5U H5 NICHICON UPW1HR47MDD CORNELL DUBILIER 715P15256JD3 KINGBRIGHT STMICROELECTRONI CS WP710A10ID5V 14 V 24 VDC D FLIP FLOP QUAD NAND GATE RELAY DRIVER SN74HC14NE4 SN74HC74 SN74ACH00N Overvoltage Protection C60 10 µF C61 100 nF C62 0.47 µF C63 1,500 pF D14 LED 13 mA 50 VDC 50 VDC 50 VDC 600 VDC 5 VDC D15 STTH102 1A 200 V P13 220 kΩ 3W VISHAY PE30L0FL224KAB P14 500 Ω 1/4 W BOURNES INC. 3362P-1-501LF R73 1 kΩ 1/4 W R74 22 Ω 1/4 W R75 10 kΩ 1/4 W R76 R77 820 Ω 4.7 Ω 1/4 W 1/2 W STACKPOLE RC12JB4R70 R78 3.3 kΩ 1/4 W R79 47 Ω 1/4 W R80 15 Ω 1W OHMITE OA150KE R81 15 Ω 1W OHMITE OA150KE R84 2 kΩ 100 W RIEDON ON SEMICONDUCTOR PF2472-2KF1 Q5 BJT 600 mA 40 VDC 110 STTH102 P2N2222AG Subcircuit Item Value/ Type Q6 IGBT DUAL COMPARATOR HEX LOGIC INVERTER U25 U26 I or P Rating 50 ADC U27 Gate Driver 200/420 mA Z13 18 V Zener 18 VDC Z14 18 V Zener 18 VDC V Rating 1200 VDC 10-20 VDC Manufacturer INTERNATIONAL RECTIFIER Manuracturer Part # FAIRCHILD TEXAS INSTRUMENTS INTERNATIONAL RECTIFIER MICRO COMMERCIAL CO MICRO COMMERCIAL CO LM393N PANASONIC EEU-FC1E101S PANASONIC EEU-FC1E101S PANASONIC EEU-FC1E101S PANASONIC EEU-FC1E101S PANASONIC EEU-FC1E101S PANASONIC EEU-FC1E101S PANASONIC 25SEP10M+TSS PANASONIC EEU-FC1E101S PANASONIC 25SEP10M+TSS PANASONIC EEU-FC1E101S PANASONIC EEU-FC1E101S NICHICON UHE2A101MPD KINGBRIGHT WP710A10ID5V KINGBRIGHT WP710A10ID5V LITTELFUSE EMERSON NETWORK POWER 0312005.HXP GE CRITICAL POWER SHHD001A3B41Z CUI INC. EMERSON NETWORK POWER EMERSON NETWORK POWER PYB10-Q48-D12 IRG7PH35UDPBF SN74HC14NE4 IR2117PBF 1N5355B-TP 1N5355B-TP On-board Power Supplies C64 100 µF C65 100 µF C66 100 µF C67 100 µF C68 100 µF C69 100 µF 10 µF, ESR = 0.06 Ω C70 C71 C72 100 µF 10 µF, ESR = 0.06 Ω C73 100 µF C74 100 µF C75 100 µF D16 LED 13 mA D17 LED 13 mA F1 Fast acting fuse 12 V Isolated DC-DC 12 V Isolated DC-DC +/-12 V Isolated DC-DC 12 V Isolated DC-DC 12 V Isolated DC-DC 5A PS1 PS2 PS3 PS4 PS5 500 mA 1.3 ADC 416 mA/Ch 500 mA 500 mA 25 VDC 25 VDC 25 VDC 25 VDC 25 VDC 25 VDC 25 VDC 25 VDC 25 VDC 25 VDC 25 VDC 100 VDC 5 VDC 5 VDC 250V 18-75 VDC 18-75 VDC 18-75 VDC 18-75 VDC 18-75 VDC 111 ASA00B36-L ASA00B36-L ASA00B36-L Subcircuit Item I or P Rating PS6 Value/ Type 12 V Isolated DC-DC R82 820 Ω 1/4 W R83 820 Ω 5 V Linear regulator 5 V Linear regulator 1/4 W Z15 Z16 500 mA 150 mA 150 mA V Rating 18-75 VDC 6-26 VDC 6-26 VDC 112 Manufacturer EMERSON NETWORK POWER Manuracturer Part # FAIRCHILD TL750L05CLPR FAIRCHILD TL750L05CLPR ASA00B36-L APPENDIX C BILL OF MATERIALS LIST 113 Table C-1 Bill of Materials. Item Qty Component Number Manufacturer Manuracturer Part # $/each Subtotal 1 2 C1,C2 PANASONIC ECO-S2EP102DX 5.02 10.04 2 1 C3 LNT2V331MSEC 15.47 15.47 3 5 715P15256JD3 1.65 8.25 4 12 K103K10X7RF5U H5 0.202 2.424 5 6 C4,C5,C6,C7,C63 C8,C10,C12,C14,C44, C45,C47,C48,C49,C50, C52,C53 C9,C11,C13,C15,C62,C 82 NICHICON CORNELL DUBILIER 0.35 2.10 6 4 0.34 1.36 7 25 0.37 9.25 8 0.31 1.24 0.44 1.76 0.94 0.94 VISHAY NICHICON UPW1HR47MDD K820J15C0GF5TH 5 C16,C19,C22,C25 C17,C18,C20,C21,C23, C24,C26,C27,C29,C30, C31,C32,C33,C34,C36, C37,C38,C40,C41,C42, C55,C57,C58,C59,C60, C61 VISHAY 4 C28,C35,C39,C43 VISHAY 9 4 C46,C51,C54,C60 PANASONIC 10 1 C56 VISHAY EEU-FC1H100 K105K20X7RF5U H5 11 9 C64,C65,C66,C67,C68, C69,C71,C73,C74 PANASONIC EEU-FC1E101S 0.34 3.06 12 2 C70,C72 PANASONIC 25SEP10M+TSS 1.56 3.12 13 1 C75 NICHICON 0.56 0.56 14 2 C76,C77 KEMET 11 22 15 1 C83 VISHAY UHE2A101MPD C4BTJBX4500ZAJ J K224K20X7RF5TH 5 16 4 CT1,CT2,CT3,CT4 LEM USA LAH 25-NP 22.88 91.52 17 4 D1,D2,D3,D4 FAIRCHILD RHRG5060 4.59 18.36 18 4 D5,D7,D9,D11 1N914BTR STTH102 0.064 0.256 19 5 D6,D8,D10,D12,D15 FAIRCHILD STMICROELECTRO NICS 0.51 2.55 20 4 D13,D14,D16,D17 KINGBRIGHT CO WP710A10ID5V 0.33 1.32 21 1 F1 LITTELFUSE 0312005.HXP 0.38 0.38 22 2 J1 FCI 10090770-S154ALF 1.02 2.04 23 1 J1 ASSMANN WSW AK532-2-R 6.08 6.08 24 2 J4,J5 MOLEX 399100102 6.92 13.84 25 2 J6,J7,J8 TE CONNECTIVITY 282858-2 1.76 3.52 TRANEX BADGER N/A 115.78 463.12 MAGNETICS INC 00K8044E026 4.36 17.44 MAGNETICS INC 00B802001 5.54 22.16 26 4 L1,L2,L3,L4 K104K10X7RF5U H5 K332K15X7RF5TL 2 VISHAY 114 Item Qty Component Number P1,P2,P3,P4,P5,P6,P7,P8 , P9,P10, P11,P12 Manufacturer Manuracturer Part # $/each Subtotal 27 12 BOURNES, INC. 3362P-1-103LF 0.81 9.72 28 1 P13 VISHAY PE30L0FL224KAB 23.9 23.9 29 1 P14 3362P-1-501LF 0.98 0.98 30 4 PS1,PS4,PS5,PS6 BOURNES, INC. EMERSON NETWORK PWR ASA00B36-L 18.31 73.24 31 1 PS2 GE CRITICAL PWR SHHD001A3B41Z 18.91 18.91 32 1 PS3 CUI INC. PYB10-Q48-D12 25.87 25.87 33 4 PT1,PT2,PT3,PT4 LV 20-P 42.35 169.4 34 4 Q1,Q2,Q3,Q4 STW88N65M5 20.9 83.6 35 1 Q5 P2N2222AG 0.38 0.38 36 1 IRG7PH35UDPBF 9.18 9.18 37 10 Q6 R1,R2,R3,R4,R5,R6,R7, R8,R80,R81 LEM STMICROELECTRO NICS ON SEMICONDUCTOR INTERNATIONAL RECTIFIER OHMITE OA150KE 2.62 26.2 38 5 R9,R12,R15,R18,R70 STACKPOLE CF14JT390R 0.08 0.4 39 4 R10,R13,R16,R19 STACKPOLE CF14JT5K60 0.08 0.32 40 4 STACKPOLE RC12JB4R70 0.56 2.24 41 17 R11,R14,R17,R20, R21,R22,R25,R26,R29, R30,R33,R34,R58,R59, R60,R61,R63,R64,R65, R69, R75 TE CONNECTIVITY 1623927-4 0.0512 0.87 42 4 STACKPOLE CF12JT100R 0.14 0.56 43 9 R24,R28,32,R36 R23,R27,R31,R35,R39, R44,R49,R54, R68 STACKPOLE 0.08 0.72 44 4 R40,R43,R48,R53 YAEGO CF14JT4K70 MFR-25FBF52133R 0.1 0.4 45 1 R41 STACKPOLE RSF2JT6K20 0.44 0.44 46 2 R42,R87 VISHAY CW00540K00JE73 1.19 2.38 47 2 R47,R52 TE CONNECTIVITY 1623782-7 0.57 1.14 48 4 R57,R66,R67,R74 YAEGO CFR-25JB-52-22R 0.1 0.4 49 1 R71 YAEGO CFR-25JB-52-560R 0.1 0.1 50 10 R37,R38,R45,R46,R50, R51,R55,R56,R72,R73 STACKPOLE CF14JT1K00 0.08 0.8 51 3 R74,R82,R83 STACKPOLE CF14JT820R 0.08 0.24 52 2 R77,R79 STACKPOLE CF14JT47R0 0.08 0.16 53 1 R78 STACKPOLE CF14JT3K30 0.08 0.08 54 1 R84 Riedon PF2472-2KF1 7.22 7.22 55 1 RELAY 1 OMRON G9EC-1 DC12 237.15 237.15 56 1 RELAY 2 PANASONIC DS2E-SL2-DC5V 6.37 6.37 57 1 SW1 JUDCO 40-2531-01 2.67 2.67 58 1 SW2 OMRON A22E-M-01 27.3 27.3 115 Item Qty Component Number Manufacturer Manuracturer Part # $/each Subtotal 38280-0112 19.64 19.64 59 1 TB1 MOLEX 60 1 U1 MICROCHIP DM240001 97.5 97.5 61 4 U2,U4,U6, U8 HCPL2531 2.21 8.84 62 4 L6387E 2.07 8.28 63 8 U3,U5,U7,U9 U10,U11,U12,U13,U14, U15,U16,U17 FAIRCHILD STMICROELECTRO NICS TEXAS INSTRUMENTS TL074CN 0.62 4.96 64 3 U18,U19,U24 LM393N 0.41 1.23 65 2 U20,U26 SN74HC14NE4 0.52 1.04 66 1 U21 SN74HC32N 0.36 0.36 67 1 U22 SN74HC74 0.36 0.36 68 1 U23 FAIRCHILD TEXAS INSTRUMENTS TEXAS INSTRUMENTS TEXAS INSTRUMENTS TEXAS INSTRUMENTS SN74ACH00N 0.36 0.36 69 1 U24 ULN2803 2.33 2.33 70 1 U27 TOSHIBA INTERNATIONAL RECTIFIER IR2117PBF 3.03 3.03 71 4 Z1,Z2,Z3,Z4 LITTELFUSE P6KE350A 0.61 2.44 1N5355B-TP Z5,Z6,Z7,Z8,Z9, Z10,Z11,Z12,Z13,Z14 MICRO COMMERCIAL CO 0.52 5.2 TL750L05CLPR 0.7 1.4 N/A 430.1 430.1 RA7031-23 94.68 94.68 72 10 73 2 Z15,Z16 74 1 75 1 FAIRCHILD ADVANCED CIRCUITS RACKMOUNTMAR T 76 1 MOLEX 38280-0112 19.64 19.64 77 11 TE CONNECTIVITY 1-390261-2 0.181 1.991 78 13 1-390261-3 0.22 2.86 79 1 TE CONNECTIVITY ON SHORE TECHNOLOGY 0.2 0.2 80 3 MILL-MAX ED18DT 917-43-103-41005000 1.89 5.67 81 1 05200101Z 1.22 1.22 82 55 LITTELFUSE KEYSTONE ELECTRONICS 5012 0.2252 12.386 83 1 ERITECH 55 55 84 2 HEAT SINK USA EGBA14212NN 10.080" Wide Extruded Aluminum Heatsink 32.76 65.52 85 4 SOUTHWIRE 0.322 1.288 86 6 SOUTHWIRE 0.116 0.696 87 1 0.076 0.076 88 8 SOUTHWIRE HARBOUR INDUSTRIES M22759/11-22-0 0.35 2.8 89 3.4 GRAINGER 2EYN8 1.42 4.7586 116 Item 90 Qty 2.4 Component Number Manufacturer Manuracturer Part # $/each Subtotal GRAINGER 1NZD4 5.92 14.05208 BURNDY YA8CLBOX 1.65 13.2 1.175 18.8 91 8 92 16 93 8 FASTENAL 1128854 0.0726 0.5808 94 24 FASTENAL 76344 0.1115 2.676 95 16 FASTENAL 1133070 0.0304 0.4864 96 8 FASTENAL 1133610 0.0172 0.1376 97 12 FASTENAL 1128879 0.0249 0.2988 98 4 FASTENAL 1128930 0.1288 0.5152 HILLMAN GROUP 99 32 FASTENAL 1136024 0.0308 0.9856 100 14 FASTENAL 33071 0.0303 0.4242 101 14 FASTENAL 0177556 0.1087 1.5218 102 24 FASTENAL 1129162 0.054 1.296 103 6 FASTENAL 76357 0.141 0.846 104 6 FASTENAL 1133618 0.0154 0.0924 TOTAL $2,367.09 117